White background with no visible content. MECHANICS OF ENGINEERING. CONTAINING STATICS AND DYNAMICS OF SOLIDS ; THE MECHANICS OF THE MATERIALS OF CONSTRUCTION, OR STRENGTH AND ELASTICITY OF BEAMS, COLUMNS, SHAFTS, ARCHES, ETC.; AND THE PRINCIPLES OF HYDRAULICS AND PNEUMATICS, WITH APPLICATIONS. FOR USE IN TECHNICAL SCHOOLS. BY IRVING P. CHURCH, C.E., PROFESSOR OF APPLIED MECHANICS, COLLEGE OF CIVIL ENGINEERING, CORNELL UNIVERSITY. EIGHTH EDITION OF "STATICS AND DYNAMICS" SEVENTH EDITION OF "MECHANICS AND MATERIALS" FIFTH EDITION OF "MECHANICS OF FLUIDS" THIRD THOUSAND. NEW YORK: JOHN WILEY & SONS, 35 EAST THIRTY STREET. 1893. A scanned page from a book titled "Mechanics of Engineering." The title is at the top, followed by a subtitle about statics and dynamics of solids. Below that, it mentions materials of construction, beams, columns, shafts, arches, and principles of hydraulics and pneumatics with applications. The text is signed off by Irving P. Church, C.E., Professor of Applied Mechanics, College of Civil Engineering, Cornell University. It also mentions the eighth edition of "Statics and Dynamics," seventh edition of "Mechanics and Materials," and fifth edition of "Mechanics of Fluids." The bottom of the page has a stamp indicating it is from the New York Public Library. THE : W YORK PUBLIC LIBRARY 702834 A ASTOR, LENOX AND TILGER FOUNDATIONS L 1954 Copyright, 1960, BY James P. Corcoran. Image of a book cover with text "HIOV WII GLISUM VASILLI" and "Disseminated by DORCHESTER & HAG, New York." Printed by PERMUS BOOKS, New York. Printed by PERMUS BOOKS, New York. PREFACE For the engineering student, pursuing the study of Applied Mechanics as part of his professional training, and not as additional mathematical culture, not only is a thoroughly systematic, clear, and consistent treatment of the subject quite essential, but one which presents the material in such a way that the student may grasp its concreteness as possible, with all the aids of the printer's and engraver's arts; and especially one which, besides showing the derivation of formulas from principles illustrated by illustrations, and lays stress on correct use of units of measurement; for without this no reliable results can be reached, and the principal object of these formulae is frustrated. With this view in mind, after a period of ten years in teaching the Mechanics of Engineering at this institution, the writer has been led to prepare the present work, in which attention is called to the following points: The diagrams are very numerous (about one to every page; an appeal to the eye is often worth a page or verbal description). The symbols for distances, angles, forces, etc., used in the algebraic work are so simple and suggestive that they are diagrams, rendering the latter full and explicit, and thus saving time and mental effort to the student. In problems in Dynamics three kinds of arrows are used to denote velocities, velocities and accelerations respectively, and thus to prevent confusion of ideas. Illustrations and examples of a practical nature, both algebraic and numerical, are of frequent occurrence. Formulas are given in their simplest form; those (homogeneous) admitting of the use of any system of units whatever for measurements of force, space, mass, and time, in numerical substitution; and those which are not so simple are given in their most direct form directed to the matter of correct numerical substitution, especially in Dynamics where time and mass, as well as force and space, are among the quantities considered. The importance, in this connection, of frequent reference to Newton's laws is clearly indicated by means employed, if also recognised, and a corresponding phrasingology adopted. The definition of force ($\vec{F}$) is made to include and illustrate Newton's law of motion; this being done in such a manner as to lead to such lengthy discussions in technical journals every few years. In the matter of "Centrifugal force," the artificial method, so commonly adopted, of regarding a particle moving uniformly in a circle as equivalent to a force acting perpendicularly to its path is shown to be one of which the "Centrifugal force" has been avoided, as being at variance with a system of Mechanics founded on Newton's laws, according to the law of which a particle moving in any other than a straight line cannot be in equilibrium. In such a system of Mechanics nothing can Iv PREFACE. be recognized as a force which is not a definite pull, push, pressure, rub, attraction or repulsion, of one body upon, or against, another. It is true that the artificial nature of the method referred to is in some textbooks fully explained in the context. (In Goodwin's Steam Engine, Knight's Mechanics, and others.) But these explanations do not mention at all, so that the student risks being led into error in attempting kindred problems by what would then seem to him correct methods. The general theorem of Work and Energy in machines is developed gradually by definite and limited steps, in preference to giving a single demonstration which, from its generality, might be too vague and abstract to be well understood by the student. In the use of the Calculus, (in the elements of which the student is supposed to have had the training usually given in technical schools by the end of his first year,) the student is often led to consider indefinite summation (except on p. 807) while the name of anti-derivative of a given function (of one variable) has been adopted for that function whose derivative, or differential coefficient, is the given function (see §356.) The signs $^{\prime}$ and $\ddot{^{\prime}}$ are used for perpendicular and parallel, respectively. In Torsten and Pleasure of Beams, the well worn and simple theory of Navier has been extended to include the graphical representation of formulae for safe loads and deflections of beams and shafts; and prominenace has been given to the methods of designing the cross-sections and rivets of built-beams and box-girders, forming the basis of the tables and rules usually given in the pocket-books of our iron and steel manufacturers. The analytical treatment of the continuous girder is not presented in the general form which belongs to it; but this is done by a method by Mohr, so greatly superior in simplicity, directness, and interest. For similar reasons the graphic of the arch of masonry is to be preferred to the analysis given by Linsley. whose insertion is chiefly a concession to the mathematical student, as are also §§ 119, 120, 264, 265, 266, 268, 264, 267, 291, and 297. The graphics of curved beams or each rise is made to precede that of the straight beam; because it is found that in many cases the graphic case of the former is then a comparatively simple matter. Hence Prof. Eddy's methods* (inserted by his kind permission) for the arch rib of hinged arches, and also that of fixed ends, are presented as special geometrical devices, intended to be but brief hints. Ed. Eddy's theorem (involving a straight girder of the same section and mode of support). Acknowledgment is also due Prof. Burr and Prof. Robinson, for their contributions to this work. Items and passages from their works (see §§304, 319, 390, and 397.) * See page 104 & §304 "Resonance" Graphical Method" by Prof. E. S. Eddy, C.E.P.H., published by J. Van Tassell New York, 1858 reprinted from Van Tassell's Magazine for 1871. PREFACE Advantage has been taken of the results of the most recent experimental investigations in Hydraulics in assigning values of the numerous coefficients necessary in this science. The researches of Messrs. Priestley and Sturz to determine the power in steam, and of Mr. Clements Herschel in testing his invention the "Venturi Water-meter," are instances in point; as also some late experiments on the transmission of pressure through pipes, by Mr. H. W. Boulton, Esq., and extensive investigations is the Hydraulics of Fire-streams and resistance of Fire-hose, p. 832. (See Transac. Am. Soc. Civ. Engg. for Nov. 1885.) In dealing with fluid tension care has been taken to use the absolute pressure as a basis, and to avoid any confusion which might arise from the occurrence of the term "negative pressure;" this precaution being specially necessary in the treatment of gaseous fluids. Though attention is given to the great length with the action of fluid motors, sufficient matter is given in treating of the mode of working of steam, gas, and hot-air engines, air-compressors, and pumping-engines, together with numerical examples, to be of considerable advantage, it is thought, however, that no attempt will be made to give a complete treatise on this subject. Special acknowledgment is due to Col. J. T. Fanning, the well-known author of "Hydraulic and Water-supply Engineering," for his consent to the use of his data respecting the frictional resistance of water in pipes, given in that work; and to Prof. C. E. Cranfield, of this university, for permission to incorporate the chapter on Retaining-Walls. References to original research in the Hydraulic Laboratory of the Civil Engineering Department at this institution will be found on pp. 694 and 720. CORNELL UNIVERSITY, ITHACA, N. Y., January, 1890. 1
TABLE OF CONTENTS.
[MECHANICS OF SOLIDS.]
PRELIMINARY CHAPTER §§ 1-10a. Definitions. Kinds of Quantity. Homogeneous Equations. Parallelogram of Forces. PART I. STATICS. CHAPTER I. STATICS OF A MATERIAL POINT. §§ 16-19. Composition and Equilibrium of Concurrent Forces. 8 CHAPTER II. PARALLEL FORCES AND THE CENTRE OF GRAVITY. §§ 50-52. Parallel Forces. 18 §§ 55-57b. Centre of Gravity. Problems. Centroidal Method. 18 CHAPTER III. STATICS OF A RIGID BODY. §§ 58-64. Couples. 37 §§ 65-69. Composition and Equilibrium of Non-concurrent Forces. 81 CHAPTER IV. STATICS OF FLEXIBLE COINS. §§ 60-68. Pulleys. Suspended Weight. Parabolic Cord. Catenary. 43 PART II. DYNAMICS. CHAPTER I. RECTILINEAR MOTION OF A MATERIAL POINT. §§ 69-75. Uniform Motion. Falling Bodies. Newton's Laws. Mass... 49 §§ 76-80. Uniformly Accelerated Motion. Graphic Representations. Variably Accelerated Motions. Impact... 54 CHAPTER II. VIRTUAL VELOCITIES. §§ 81-89. Definitions and Propositions. Connecting Rods. Problems... 67 VII viii CONTENTS. CHAPTER III. CURVILINEAR MOTION OF A MATERIAL POINT. PAGE 70-74. Composition of Motions, of Velocities, etc. General Formulae. 73 Numerical Substitution. Centralial and Central Forces. Simple Pendulum. Projectiles. Rele- tive Motion. 77 CHAPTER IV. MOMENT OF INERTIA. 85-94. Plane Figures. Rigid Bodies. Reduction Formulae. The Rectangle, Triangle, etc. Compound Plane Figure. Polar Axis. 91 Roots. This Problem. Geometric Solids. 96 Numerical Substitution. Ellipsoid of Inertia. 102 CHAPTER V. DYNAMICS OF A ROUGH BOAT. 106-118. Translation. Rotation about a Fixed Axis. Centre of Percussion. 106 Torsion Balance. Compound Pendulum. The Fly-wheel. 116 Uniform Rotation. " Centrifugal Acceleration." Free Roll. 135 Rolling Motion. Parallel Rod of Locomotive. 130 CHAPTER VI. WORK, ENERGY, AND POWER. 137-146. Work. Power. Horse-power. Kinetic Energy. 138 Steam-hammer. Pile driving. Inelastic Impact. 138 Rotary Motion. Equivalent Systems of Forces. Any Machine. 148 Work and Kinetic Energy in a Moving Machine of Rigid Parts. 147 Power of Motor. Potential Energy. Heat, etc. Dy- namic Work in Machines. Boat steering. 158 CHAPTER VII. FRICTION. 158-164. Sliding Friction. Its Lawe. Best Lever. 164 Axis-friction. Friction Wheela. Pivota. Boiting. 175 Transmission of Power by Boiting. 175 Rolling Friction. Brakes Engine-friction Anoma- lies in Friction. Rigidity of Cordsa. 196 CONTENTS. ix PART III. STRENGTH OF MATERIALS. (OR MECHANICS OF MATERIALS.) CHAPTER I. ELEMENTARY STRESSES AND STRAINS. § 178-182. Stress and Strain ; of Two Kinds. Oblique Section of Rod in Tension and Compression. . . . . . . . . . . . . . . . 198 § 183a-190. Hooke's Law. Elasticity. Safe Limiit. Elastic Limit. Rupture. Modulus of Elasticity. Isotropy. Resilience. Internal Stress. Temperature Effects. . . . . . . . . . . . . . . . 301 TENSION. § 191-195. Hooke's Law by Experiment. Strain Diagrams. Lateral Contraction. Modulus of Tenacity. 307 § 196-199. Resilience of Stretched Prime. Load Applied Suddenlly. Load Applied to One Cown Work. Solid of Uniform Strength. Tensileum Rupture. 313 COMPRESSION OF SHORT BLOCKS. § 200-202. Short and Long Columns. Remarks on Crushing... 218 EXAMPLES IN TENSION AND COMPRESSION. § 203-205. Tables. Examples. Factor of Safety. Practical Notes. 290 SHREARING. § 207-313. Rivets. Shearing Distortion. Table. Punching. Examples... 325 CHAPTER II. TOBSSION. § 314-320. Angle and Moment of Torsion. Torsional Strength, Stiffness, and Resilience. Non-Circular Shafts... 338 § 321-328 Transmission of Power. Autographic Testing Machine. Examples... 388 CHAPTER III. FLEXURE OF HOMOGENEOUS PRISMS UNDER PERPENDICULAR FORCES IN ONE PLANE. § 324-332a The Common Theory of Elastic Forces. Neutral Axis : The "Moment" and "Moment" Flexural Strength and Strain, Radius of Curvature... 344 ELASTIC CURVES. § 333-336 Single Central Load ; at Rest, and Applied Suddenly. Eccentric Load. Uniform Load. Cantilever... 352 X CONTENTS. **SAFE LOADS.** §§ 389-340. Maximum Moment. Shear = x-Derivative of the Moment. Simple Beams With Various Loads. Comparative Strength and Stiffness of Rectangular Beams. 262 §§ 347-358. Moment of Inertia of Rectangular Beams. Cantilevers. Tables. Numerical Examples. 273 **SHIELDING STRESSES IN FLEXURE.** §§ 353-357. Shearing Stress Parallel to Neutral Surface; and in Cross Section. Web of I-Beam. Biretting of Built Beams. 284 **SPECIAL PROBLEMS IN FLEXURE.** §§ 355-365. Designing Sections of Built Beams. Moving Loads. Special Cases of Quiescent Loads. Hydrostatic Load. 295 §§ 366-270. Derivatives of Ordinates of Elastic Curve. Weight Pulling on Beam. Crane Shaft. Other Shafts. 310 CHAPTER IV. FLEXURE: CONTINUED. **CONTINUOUS BEAMS.** §§ 271-276. Analytical Treatment of Symmetrical Cases of Beams on Three Supports; Also, Built-in (see § 361.). 390 THE DANGEROUS SECTION IN NON-PRISSMATIC BEAMS. §§ 277-278. Double Truncated Wedge, Pyramid, and Cone. 391 §§ 280-281. Non-Priismatic Beams of Uniform Strength. Parabolic and Wedge-Shaped Beams, L-Beams, Elliptical Beam and Cantilevers. 395 DEFORMATION OF BEAMS OF UNIFORM STRENGTH. §§ 290-293. Parabolic and Wedge-Shaped Beams. Special Problems... 843 CHAPTER V. FLEXURE OF PRISMATIC BEAMS UNDER OBLIQUE FORCES. §§ 294-301. Gravity and Neutral Axes. Shear, Thrust, and Stress-Couple. Strength and Stiffness of Oblique Cantilever. Hooks. Crane... 347 CONTENTS. xii CHAPTER VI. FLEXURE OF LONG COLUMNS. §§ 305-306. End-Conditions. Euler's Formula. "Incompetent Flexure." Hodgkinson's Formulae. Rankine's Formulae. §§ 307-310. Radius of Gyration. Built Columns. Trussed Girder. Buckling of Web-Plates. Examples. §§ 311-312. Inverted Catenary. Parabolic Arch. Circular Arch. Transformed Catenary as Arch. CHAPTER VII. LINEAR ARCHES (OF BLOCKWORK). §§ 315-328. CHAPTER VIII. ELEMENTS OF GRAPHICAL STATICS §§ 324-326. Force Polygons. Concurrent and Non-Concurrent Forces in a Plane. Force Diagrams. Equilibrium Polygons. §§ 327-332. Constant Stress in a Beam. Elastic Stresses in Roof Truss. Bow's Notation. The Special Equilibrium Polygon. CHAPTER IX. GRAPHICAL STATICS OF VERTICAL FORCES. §§ 333-336. Joined Rods. Centre of Gravity. §§ 337-343. Useful Relations Between Force Diagrams and Their Equilibrium Polygons. CHAPTER X. RIGHT ARCHES OF MASONRY. §§ 344-352. Definition. Mortar and Friction. Pressure in Joints. Conditions of Safe Equilibrium. True Linear Arch. §§ 353-357. Arrangement of Details. §§ 358-360. Graphical Treatment of Arch. Symmetrical and Unsymmetrical Cases. CHAPTER XI. ARCH-RIBS. §§ 361-374. Mode of Support. Special Equilibrium Polygon and its Force Diagram. Change in Angle Between Rib and Tangentie. Displacement of Axis Point on Rib. §§ 375a-378a. Graphical Arithmetic, Summation of Products, Moment of Inertia by Graphics, Classification of Arch-Ribs. §§ 379-388. Profiling of Ribs, Braced Ribs, Hinged Ends, and of Fixed Ends. Stress Diagrams Temperature Stresses. Braced Arches. 401 xii CONTENTS. HORIZONTAL STRAIGHT GIRDES. §§ 880-890. Prismatic Girder Supported at the Extremities, also with Ends Fixed Horizontally. 479 CHAPTER XIII. GRAPHICS OF CONTINUOUS GIRDES. §§ 391-397. The Elastic Curve an Equilibrium Polygon. Mohr's Theorem. Example and Numerical Case. End-Tangents. Re-arrangement of Moment-Area. 484 §§ 398-406. Positive Moment-Areas ; Amount and Gravity Verticals of Moments. The Moment Curve in Any Given Case, Moment Curves, Shears, and Reactions. Numerical Example in Detail. Remarks 497 [CONTENTS OF "MECHANICS OF FLUIDS."] PART IV.—HYDRAULICS. CHAPTER I.—DEFINITIONS. FLUID PRESSURE. HYDRO-STATICS. 406–417. Perfect fluid. Elasticity of Fluids. Principle of "Equal Transmission of Pressure." Non-planar Pistons... 515 418–437. Hydraulic Press. Free Surface of Liquid. Barometers and Manometers. Safety-valves. Strength of Thin Hollow Cylinders against Bursting and Collapse... 535 CHAPTER II.—HYDROSTATICOS CONTINUED. PRESSURE OF LIQUIDS IN TANKS AND RESERVOIRS. 438–484. Liquid in Motion, but in Relative Equilibrium. Theorems on the Surface of Pressure of Rectangles, Triangles, etc.... 540 485–444. Stability of Rectangular and Trapezoidal Walls against Water Pressure. High Massory Dams. Propped Quay Bridge. Effect of a Dam. Water Pressure on both Sides of a Gate... 554 CHAPTER III.—EARTH PRESSURE AND RETAINING WALLS. 445–465. Angle of Repose. Wedge of Maximum Thrust. Geometric Stability of Retaining Walls. Results of Experience... 579 CHAPTER IV.—HYDROSTATICS CONTINUED. IMMERSION AND FLotation. 466–460. Buoyant Effort. Examples of Immersion. Specific Gravity. Equilibrium of Flotation. Hydrometer... 595 461–465. Depth of Flotation. Draught and Angular Stability of Ships. The Metacentre... 599 zill xiv CONTENTS. CHAPTER V.—HYDROSTATICS CONTINUED. GASEOUS FLUIDS. 406-478 Thermometers. Absolute Temperature. Gases and Ve- pers. Critical Temperature. Law of Charles. Closed Air-manometer. Mariotte's Law. Mixture of Gases. Boyle's Law. Pressure of a Gas. 304 Work Done in Steam-engine Cylinders. Expanding Steam. Graphic Representation of Change of State of a Gas. Heat of Compressibility. 305 Hot-air Engines. Gas-engines. Heat-efficiency. Duty of Pumping-engines. Buoyant Effort of the Atmosphere 324 CHAPTER VI.—HYDRODYNAMICS BEGINN. STEADY FLOW OF LIQUIDS THROUGH PIPES AND ORIFICES. 486-495 Phenomena of a "Steady Flow." Bernoulli's Theorem for Steady Flow without Friction, and Applications. Circular Orifice. Velocity of Liquid through Orifice 346 496-500 Rounded Orifice. Various Problems Involving Flow through Orifice. Jet from Force-pump. Velocity and Pressure at Orifice. Velocity and Pressure from Vessel in Motion. Barker's Mill 368 501-506 Efflux from Rectangular and Triangular Orifices. Pen- etration of Liquid through Orifices and Nozzles. Contraction, etc. Overall Weir. Experiments of Francis, Freely and Stearns, and Bazin. Short Pipes or Tubes 373 506-513 Correlation of Vessels, Valves, etc., with Field Friction. Freude's Experiments. Bernoulli's Theorem with Friction. Hydraulic Radius. Loss of Head. Prob- lems involving Friction Heads in Pipes. Accumu- lation of Water in Pipes 378 513a-518 Loss of Head in Orifices and Short Pipes. Coefficient of Friction of Water in Pipes. Fanning's Table. Bernoulli's Theorem with Friction in Pipes 708 519-526 Chizky's Formula. Fire-engine Hose. Pressure-energy. Losses of Head due to Sudden Enlargement of be- hind the nozzle, etc. Discharge in Pipe. Ven- turi Water-meter 714 527-536 Sudden Diminution of Section. Losses of Head due to Elbows, Bends, Valve-gates, and Throttle-valves. Effect on Velocity and Pressure of Liquid. Siphons. Branching Pipes. Time of Emptying Vessels of Various Forma; Prism, Wedges Pyra- mids, etc., etc., etc., etc., etc., etc., etc., etc., etc., Volumes of Irregular Form using Simpson's Rule 727 CONTENTS. XIV CHAPTER VII.—HYDRODYNAMICS, CONTINUED; STEADY FLOW OF WATER IN OPEN CHANNELS. §§ 530-543a. Nomenclature. Velocity Measurements and Instruments for the same Ritchie-Haskell Direction Current Meter. Measurement of Velocity with Depth Pitot's Tube. Hydrometer. Gauging Streams. Chezy's Formula for Uniform Motion in Open Channel. Experimental Methods. 749 §§ 543b-547. Kutter's Formula. Sections of Least Resistance. Trapezoidal Section of Given Side Slope and Minimum Friction Coefficient. Bends. Formula Introducing Depth at End Sections. Backwater. 759 CHAPTER VIII.—DYNAMICS OF GASEOUS FLUIDS. §§ 548-556. Theorem for Steady Flow of Gas without Friction. Theoretical Relations between Pressure and Velocity by Water formula; with Isothermal Expansion; with Adiabatic Expansion. Maximum Flow of Weight. Experimental Considerations in Long and Short Pipes. Force Considering Velocity of Apparatus. 778 §§ 557-561a. Transmission of Compressed Air through Long Pipes. Experiments in St. Gotthard Tunnel. Pipes of Variable Diameter. The Piping of Natural Gas. 786 CHAPTER IX.—IMPULSE AND RESISTANCE OF FLUIDS §§ 562-566. Reaction of a Jet of Liquid. Impulse of Jet on Curved Vanes, Fixed and in Motion. Pitot's Tube. The California "Hurdy-gurdy". Impulse on Plates Flows Moving in a Fluid. Forces on Curves of Fluid. §§ 570-575. Wind-pressure. Smitsonian Scale. Mechanics of the Sail-boat. Resistance of Still Water to Immersed Solid Bodies; to Bodies in Vertical Plane; to Bodies in Vertical Plane. Robinson's Cup anemometer. Resistance of Ships. Transporting Power of a Current. Fire-streams, Hose-friction, etc. 818 --- MECHANICS OF ENGINEERING. PRELIMINARY CHAPTER. 1. Mechanics treats of the mutual actions and relative motions of material bodies, solid, liquid, and gaseous; and by Mechanics of Engineering is meant a presentment of those principles of pure mechanics, and their applications, which are of special service in engineering problems. 2. Kinds of Quantity.—Mechanics involves the following fundamental kinds of quantity: *Space*, of one, two, or three dimensions, i.e., length, surface, or volume, respectively; *time*, which needs no definition here; *force* and *mass*, as defined below; and *absolute* numbers, whose values are independent of arbitrary units, as, for example, a ratio. 3. Force.—A force is one of a pair of equal, opposite, and simultaneous actions between two bodies, by which the state* of their motions is altered or a change of form in the bodies themselves is effected. Pressure, attraction, repulsion, and traction are instances in point. Similarity sensation conveys the idea that a beam-balance is a balance only in absolute measure of it, a beam-balance only a relative measure. In accordance with Newton's third law of motion, that action and reaction are equal, opposite, and simultaneous, forces always occur in pairs; thus, if a pressure of 40 lbs. exists between bodies $A$ and $B$, if $A$ is considered by itself (i.e., "free"), apart from all other bodies whose actions upon it are called forces, among these forces will be one of 40 lbs. directed from $B$ toward $A$. Similarly, if $B$ is under consideration, a force *The state of motion of a small body under the action of no force, or of balanced forces acting on it in different directions, remains in a right line.* is directed from $A$ toward $B$. If the motion is altered from this, the fact is due to the action of an unbalanced force (§ 54). 2 **MECHANICS OF ENGINEERING.** of 40 lbs. directed from $A$ toward $B$ takes its place among the forces acting on $B$. This is the interpretation of Newton's third law. In conceiving of a force as applied at a certain point of a body it is useful to imagine one end of an inponderable spiral spring in a state of compression (or tension) as attached at the given point, the axis of the spring having the given direction of the force. 4. **Mass** is the quantity of matter in a body. The masses of several bodies being proportional to their weights at the same locality on the earth's surface, in physics the weight is taken as the mass, but in practical engineering another mode is used for measuring it (as explained in a subsequent chapter), viz.: the mass of a body is equal to its weight divided by the acceleration of gravity in the locality where the weight is taken, or, symbolically, $m = \frac{w}{g}$. This quotient is a constant quantity, he should be since the mass of a body is invariable wherever the body be carried. 5. **Derived Quantities**—All kinds of quantity besides the fundamental ones just mentioned are compounds of the latter, formed by multiplication or division, such as velocity, acceleration, momentum, work, energy, moment, power, and force-distribution. Some of these are merely names given for convenience to certain combinations of factors which come together naturally under certain physical principles, but as a result of common algebraic transformation. 6. **Homogeneous Equations** are those of such a form that they are true for any arbitrary system of units, and in which all terms combined by algebraic addition are of the same kind. Thus, the equation $s = \frac{gt^2}{2}$ (in which $g =$ the acceleration of gravity and $t$ the time of vertical fall of a body in vacuo, from rest) will give the distance fallen through $s$, whatever units be adopted for measuring time and distance. But if for PRELIMINARY CHAPTER. 3 g we write the numerical value 32.2, which it assumes when time is measured in seconds and distance in feet, the equation $s = 16.1f$ is true for those units alone, and the equation is not of homogeneous form. Algebraic combination of homogeneous equations should always produce homogeneous equations; if not, some error has been made in the algebraic work. If any equation derived or proposed for practical use is not homogeneous, an explicit statement should be made in the context as to the proper units to be employed. 7. Heaviness.—By heaviness of a substance is meant the weight of a cubic unit of the substance. E.g. the heaviness of fresh water is 62.5, in case the unit of force is the pound, and the foot the unit of space; i.e., a cubic foot of fresh water weighs 62.5 lbs. In case the substance is not uniform in composition, the heaviness varies from point to point. If the weight of a homogeneous body be denoted by $G$, its volume by $V$, and the heaviness of its substance by $\gamma$, then $G = \gamma V$. WEIGHT IN POUNDS OF A CUBIC FOOT (i.e., THE HEAVINESS) OF VARIOUS MATERIALS
MATERIAL WEIGHT IN POUNDS OF A CUBIC FOOT
Anthracite, solid. 100 Masonry, dry rubble. 138
" broken" " dressed" (grain or).
Brick, common hard. 135 Limestone. 145
" soft." 100 Mortar. 100
Brick, common. 113 Petroleum. 55
Cement. Snow. 7
Earth, loose. 73 wet. .15 to .50
" as mud." 108 Steel. 490
Graupel. .964 to .974 " dressed" (grain or). .38 to .40
Ice, Limestone. .38 to .40
" frozen." 68 Water, fresh. .92 to .95
Iron, cast. 430 " snow." .44 to .46
" wrought." 480
8. Specific Gravity is the ratio of the heaviness of a material to that of water, and is therefore an abstract number. 9. A Material Point is a solid body, or small particle, whose dimensions are practically nothing, compared with its range of motion. 4 MECHANICS OF ENGINEERING. 10. A Rigid Body is a solid, whose distortion or change of form under any system of forces to be brought upon it in practice is, for certain purposes, inenensible. 11. Equilibrium.—When a system of forces applied to a body produces the same effect as if no force acted, so far as the state of motion of the body is concerned, they are said to be balanced or to be in equilibrium. [If no force acts on a material point it remains at rest if already at rest; but if already in motion it continues in motion, and uniformly (equal spaces in equal times), in a right line in direction of its original motion. See § 54.] 12. Division of the Subject.—Statics will treat of bodies at rest, i.e., of balanced forces or equilibrium; dynamics, of bodies in motion; strength of materials will treat of the effect of forces in distorting bodies; hydraulics, of the mechanics of liquids and gases (thus including pneumatica). 13. Parallelogram of Forces.—Duchayla's Proof. To fully determine a force we must have given its amount, its direction, and its point of application in the body. It is generally denoted in diagrams by an arrow. It is a matter of experience that between two points situated at some distance apart, any other may be chosen in the line of action of the force. This is called the transmissibility of force; i.e., so far as the state of motion of the body is concerned, a force may be applied anywhere in its line of action. The Resulant of two forces (called its components) applied at a point of a body is a single force applied at the same point, which will replace them. To prove that this resultant is given in amount and position by the sum total of the forces formed on the line given for their lines of action laid off to some scale, so many pounds to the inch, say). Duchayla's method requires four postulates, viz.: (1) the resultant of two forces must lie in the same plane with them; (2) the resultant of two equal forces must bisect the angle between them; (3) if one of the two forces be increased, the angle between the other force and the resultant will be greater than before; and (4) the transmissibility of forces, already mentioned. Granting these, we proceed as follows (Fig. 1): Given the two forces $P$ and $Q=$ PRELIMINARY CHAPTER. 5 $P' + P'' (P' \text{ and } P'')$ being each equal to $P$, so that $Q = 2P$, applied at $O$. Draw the parallelograms $OB$ and $AD$. $OD$ will also be a parallelogram. By postulate (2), since $OB$ is a rhombus, $P'$ and $P''$ at $O$ may be replaced by a single force $R'$ acting through $A$. Transit $R'$ to $B$ and replace $B$ by $D$. Transit $P'$ to $B$ and replace $B$ by $D$. Similarly $P$ and $P''$ at $A$, may be replaced by a single force $R'$ passing through $D$; transmit it there and resolve it into $P$ and $P''$. $P'$ is already at $D$. Hence $P$ and $P' + P''$, acting at $D$, are equivalent to $P$ and $P' + P''$ acting at $O$, in their respective directions. Therefore the resultant of $P$ and $P' + P''$ must lie in the line $OD$, the diagonal of the parallelogram formed on $P$ and $Q = 2P$ at $O$. Similarly Diagram showing forces P', P'', P, P', P'' acting on points O, B, D, A, C respectively. Fig. 1 this may be proved (that the diagonal gives the direction of the resultant) for any two forces $P$ and $mP'$, and for any two forces $nP$ and $mP$, $m$ and $\alpha$ being two whole numbers, i.e., for any two commensurable forces. When the forces are incommensurable (Fig. 2), $P$ and $Q$ being the given forces, we may use a reductio ad absurdum, thus: Form the parallelogram $OPQ$ on $P$ and $Q$ applied at $O$. Suppose for an instant that this is the diagonal of a parallelogram, but not so long as it is parallel to $\overline{OD}$, but some other direction, as $\overline{OD}$. Note the intersection of $\overline{H}$, and draw $\overline{HO}$ parallel to $\overline{DB}$. Divide $\overline{P}$ into equal parts, each less than $\overline{HD}$; then in laying off parts equal to these from $\overline{O}$ along $\overline{OB}$, a point of division will come at some point $\overline{F}$ between $\overline{C}$ and $\overline{B}$. Complete the parallelogram OFEG. The force $\overline{Q'} = \overline{OF}$ is commensurable with $\overline{F}$, and hence their 6 MECHANICS OF ENGINEERING. resultant acts along $OE$. Now $Q$ is greater than $Q'$, while $R$ makes a less angle with $P$ than $OE$, which is contrary to postulate (3); therefore $R$ cannot lie outside of the line $OD$. Q. E. D. It still remains to prove that the resultant is represented in amount, as well as position, by the diagonal. OD (Fig. 5) is the direction of $R$ the resultant of $P$ and $Q$; required its amount. If $R'$ be a force equal and opposite to $R$ it will balance $P$ and $Q$; i.e., the resultant of $P$ and $P'$ being equal to $Q$, and also the resultant of $Q$ and $Q'$ being equal to $Q$. We can therefore determine $R'$ by drawing $BA$ parallel to $DO$ to intersect $QQ'$ prolonged in $\mathcal{A}$; and then complete the parallelogram on $\mathcal{BA}$ and $\mathcal{BO}$. Since $\mathcal{OFAB}$ is a parallelogram $\mathcal{R'}$ must = $\mathcal{EA}$, and since $\mathcal{OARD}$ is a parallelogram $\mathcal{OA} = \mathcal{OD}$; therefore $\mathcal{R'} = \mathcal{OD}$ and also $\mathcal{R} = \mathcal{OD}$. Q. E. D. Corollary.--The resultant of three forces applied at the same point is the diagonal of the parallelopiped formed on the three forces. 14. Concurrent forces are those whose lines of action intersect in a common point, while non-concurrent forces are those which do not so intersect; results obtained for a system of concurrent forces are really derivable, as particular cases, from those pertaining to a system of non-concurrent forces. 15. Resultant.--A single force, the action of which, as regards the state of motion of the body acted on, is said to be equivalent to that of a number of forces forming a system, is said to be the Resultant of that system, and may replace the system; and conversely a force which is equal and opposite to the resultant of a system is called the Anticorrespondent, or, in other words, when it is combined with that system together it results a system in equilibrium; this (ideal) force is called the Anti-resultant. In general, as will be seen, a given system of forces can al- * R must = OP; for if P > O < OP the diagonal formed on R' and P cannot take the direction of CO prolonged. 6 PRELIMINARY CHAP. II. 7 ways be replaced by two single forces, but these two can be combined into a single resultant only in particular cases. 15a. Equivalent Systems are those which may be replaced by the same set of two single forces—or, in other words, those which have the same effect, as to state of motion, upon the given body. 15b. Formula.—If in Fig. 8 the force $P$ and $Q$ and the angle $\alpha = P O Q$ are given, we have, for the resultant, $$R = OD = \sqrt{P^2 + Q^2 + 2 PQ \cos \alpha}$$ (If $\alpha$ is > 90° its cosine is negative). In general, given any three parts of either plane triangle $O D Q$, or $O D B$, the other three may be obtained by ordinary trigonometry. Evidently if $\alpha = 0$, $R = P + Q$; if $\alpha = 180°$, $R = P - Q$; and if $\alpha = 90°$, $R = \sqrt{P^2 + Q^2}$. 15c. Varieties of Forces.—Great care should be used in deciding what may properly be called forces. The latter may be divided into actions by contact, and actions at a distance. If pressure exists between two bodies so that one body tends to move in a certain direction, this is an action by contact (or thrust), or compressive action, of one against the other constitutes a force, whose direction is normal to the tangent plane at any point of contact (a manner of expressing the fact that the force acts perpendicularly to the surface of contact). The rest may be tangential actions or friction. (If the bodies really form a continuous substance at the surface considered, these tangential actions are called shearing forces.) Again, when a rod or wire is subjected to tension, any portion of it will tend to move in a certain direction; this is an action by contact (or thrust), or tensile action, of one against the other; the ability to do this depends on the property of cohesion. The foregoing are examples of actions by contact. Actions at a distance are classified in the mysterious attractions, or reactions, observable in the phenomena of gravitation, electricity, and magnetism, where the bodies concerned are not necessarily in contact. By the term weight we shall always mean the force of the earth's attraction on the body in question, and not the amount of matter in it. (Note—in some common phrases such as "The tremendous force" of a heavy body to push another," "The tremendous weight" of a heavy body to strike energy (as explained in Chap. VII.). The mere fact that a body is in motion, whatever its mass and velocity, does not imply that it has been acted upon by a force; for at any point in the path of a cannon ball through the air, the only forces acting on it are the resistance of the air and the attraction of the earth, the latter having a vertical and downward direction.) A diagram showing two forces P and Q acting on a body O. PART I—STATICS. CHAPTER I. STATICS OF A MATERIAL POINT. 18. Composition of Concurrent Forces.—A system of forces acting on a material point is necessarily composed of concurrent forces. CASE I—All the forces in one Plane. Let $O$ be the material point, the common point of application of all the forces; $P_x$, $P_y$, etc., the given forces, making angles $\alpha_x$, $\alpha_y$, etc., with the axis $X$. By the parallelogram of forces $P$, may be resolved into and replaced by its components, $P_x \cos \alpha_x$ acting along $X$, and $P_y \sin \alpha_y$ along $Y$. Similarly all the remaining forces may be replaced by their $X$ and $Y$ components. We have now a new system, the equivalent of that first given, consisting of a set of $X$ forces having their lines of action parallel (axis $X$) and a set of $Y$ forces all acting in the line $Y$. The amount of the $X$ forces being their algebraic sum (denoted by $\Sigma X$) (since they have the same line of application) we have $$\Sigma X = P_x \cos \alpha_x + P_y \cos \alpha_y + \text{etc.} = \Sigma(P \cos \alpha),$$ and similarly $$\Sigma Y = P_x \sin \alpha_x + P_y \sin \alpha_y + \text{etc.} = \Sigma(P \sin \alpha).$$ These two forces, $\Sigma X$ and $\Sigma Y$, may be combined by the parallelogram of forces, giving $R = \sqrt{\Sigma X^2 + \Sigma Y^2}$ as the single resultant of the whole system, and its direction is determined by the angle $\alpha$: thus, $\tan \alpha = \frac{\Sigma Y}{\Sigma X}$; see Fig. 5. For equilibrium to exist, $R$ must = 0, which requires, separately, Fig. 5. STATICS OF A MATERIAL POINT. 9 $\mathbf{X} = 0$, and $\mathbf{Y} = 0$ (for the two squares ($\mathbf{XX}$) and ($\mathbf{YY}$) can neither of them be negative quantities). Case II. The forces having any direction in space, but all applied at $O$, may be resolved into components. Let $P_x$, $P_y$, etc., be the given forces, $P_z$ making the angles $\alpha_x$, $\beta_x$, and $\gamma_x$ respectively, with three arbitrary axes, $X$, $Y$, and $Z$ (Fig. 6), at right angles to each other and intersecting at $O$, the origin. Similarly let $\alpha_x$, $\beta_x$, $\gamma_x$ be the angles made by $P_z$ with these axes, and so on for all the forces. By the parallelepiped of forces, $P_z$ may be replaced by its components: $$X = P_x \cos \alpha_x, Y = P_y \cos \beta_y, Z = P_z \cos \gamma_z;$$ similarly for all the forces, so that the entire system is now replaced by the three forces, $$\mathbf{XX} = P_x \cos \alpha_x + P_z \cos \gamma_z + \text{etc};$$ $$\mathbf{YY} = P_y \cos \beta_y + P_z \cos \gamma_z + \text{etc};$$ $$\mathbf{ZZ} = P_z \cos \gamma_z + P_z \cos \gamma_z + \text{etc};$$ and finally by the single resultant $$R = (\mathbf{XX})^2 + (\mathbf{YY})^2 + (\mathbf{ZZ})^2.$$ Therefore, for equilibrium we must have separately, $$\mathbf{XX} = 0, \mathbf{YY} = 0, \text{and } \mathbf{ZZ} = 0.$$ $Fz$'s position may be determined by its direction cosines, viz., $$\cos \alpha = \frac{\mathbf{XX}}{R}, \cos \beta = \frac{\mathbf{YY}}{R}, \cos \gamma = \frac{\mathbf{ZZ}}{R}.$$ 17. Conditions of Equilibrium.—Evidently, in dealing with a system of concurrent forces, it would be a simple matter to MECHANICS OF ENGINEERING. replace any two of the forces by their resultant (diagonal formed on them), then to combine this resultant with a third force, and so on until all three forces had been combined, the last resultant being the resultant of the whole system. The foregoing treatment, however, is useful in showing that for equilibrium of concurrent forces in a plane only two conditions are necessary, viz., $\Sigma X = 0$ and $\Sigma Y = 0$; while in space there are three, $\Sigma X = 0$, $\Sigma Y = 0$, and $\Sigma Z = 0$. In Case I, then, we have conditions enough for determining two unknown quantities; in Case II, three. 18. Problems involving equilibrium of concurrent forces. A rigid body in equilibrium under no more than three forces may be treated as a material point, since the (two or) three forces are necessarily concurrent.* Problem 1.—A body weighing $G$ lbs. rests on a horizontal table; required the pressure between it and the table. Fig. 8. Consider the body free, i.e., conceive all other bodies removed (the table in this instance), being replaced by the weight of the body itself. Let us suppose that resting the axis $Y$ vertical and positive upward, and not assuming in advance either the amount or direction of $N$, the pressure of the table against the body, but knowing that $G$, the action of the earth on the body, is vertical and downward, we have here a system of concurrent forces in equilibrium, in which the $X$ and $Y$ components of $G$ are known (being 0 and $-G$ respectively), while those, $X_4$ and $X_5$, of $N$ are unknown. Putting $\Sigma X = 0$, we have $X_4 + G = 0$, i.e., $X_4 = -G$; or the vertical component of $N$, i.e., $N$ itself, is positive (upward in this case), and is numerically equal to $G$. Problem 2.—Fig. 9. A body of weight $G$ lbs. is moving in a straight line over a rough horizontal table with a uniform velocity $v$ (feet per second) to the right. The tension in an oblique cord by which it is pulled is given, and is $P$ (lbs.), * Three parallel forces form an exception; see §§20, 21, etc. A diagram showing a body moving horizontally on a rough surface with a tension force applied at an angle. 16 STATICS OF A MATERIAL POINT. 11 which remains constant, the cord making a given angle of elevation, $\alpha$, with the path of the body. Required the vertical pressure $N$ (lbs.) of the table, and also its horizontal action $F$ (friction) (lbs.) against the body. Referring by anticipation to Newton's first law of motion, viz., a material point acted on by no force or by balanced forces is either at rest or moving uniformly in a straight line, we see that this problem is a case of balanced forces, i.e., of equilibrium. Since there are only two unknown quantities, $N$ and $F$, we may determine them by the two cases of Case I., taking the axes $X$ and $Y$ as before, and considering the direction of $N$ as well as its amount to be determined by the analysis. As $F$ must evidently point toward the left, treat it as negative in summing the $X$ components; the analysis, therefore, can be expected to give only its numerical value. $$2X = 0 \text{ gives } P\cos\alpha - F = 0.\quad\therefore F = P\cos\alpha.$$ $$2Y = 0 \text{ gives } N + P\sin\alpha - G = 0.\quad\therefore N = G - P\sin\alpha.$$ $\therefore N$ is upward or downward according as $G >$ or $< P\sin\alpha$. For $N$ to be a downward pressure upon the body would require the surface of the table to be above it. The ratio of the frictional to the normal force which produces it can now be obtained, and is called the coefficient of friction. It may vary slightly with the velocity. This problem may be looked upon as arising from an experiment made to determine the coefficient of friction between the given surfaces at the given uniform velocity. **19. The Free-Body Method.** — The foregoing rather laborious solutions of very simple problems have been made such to illustrate what may be called the free-body method of treating any problem involving a body acted on by a system of forces. It consists in conceiving the body isolated from all others which act on it in any way, those actions being introduced as so many forces, known or unknown, in amount and position. The system of forces thus formed may be made to yield certain equa- A diagram showing a body at rest on a table with forces acting on it. Fig. 2 12 MECHANICS OF ENGINEERING. tions, whose character and number depend on circumstances, such as the behavior of the body, whether the forces are confined to a plane or not, etc., and which are therefore theoretically available for determining the equal number of unknown quantities, whereas these forces may vary at times or abstract numbers. Of course in some instances the unknown quantities may enter these equations with such high powers that the elimination may be impossible; but this is a matter of algebra, not of mechanics. (Addendum to § 60 of page 68.) Numerical Example.—A set of light screws lies on a horizontal plane. A cannon-ball, with the object of determining its velocity, and also the rate of change (or negative acceleration) of that velocity, as due to the resistance of the air. By electrical connection the time of passing each screen is noted, and the intervals of time are given in this diagram for four of the screens, $A$, $B$, $C$, and $D$.
$A$ $B$ $C$ $D$
Time (sec.) 100 100 100 100
Velocity (ft./sec.) 0.0631 sec. 0.0632 sec. 0.045 sec.
From these data it is required to compute, as nearly as the circumstances allow, the average velocity (negative) of the ball at various points (the ball moves from left to right). Solution.—In passing from $A$ to $B$ the ball has an average velocity of 1010 ft. per second at point 1 ; and similarly at point 2 ; and so by the time of passage, 0.063 second. Similarly we find the average velocity between $B$ and $C$ to be 1584 ft. per second, and that between $C$ and $D$ to be 1584 ft. per second. As the velocity is not changing very rapidly, we may claim that the ball actually possesses the velocity $v_1 = 1610$ ft. per second at the point 1, midway between $A$ and $B$, or very near that point ; and similarly the velocity $v_2 = 1584$ ft. per second at point 2, midway between $B$ and $C$ ; and $v_3 = 1584$ ft. per second at point 3, midway between $C$ and $D$. Hence the total gain of velocity from 1 to 2 is $\frac{1}{2} \times 0.063 \times (1610 - 1584)$ feet per second, or about 7 feet per second ; and from 2 to 3 is one half of the 0.063 second plus one half of the 0.063 second, i.e., 0.063 second. Therefore an approximate value for the average acceleration between points 1 and 2 is found to be $\frac{7}{(0.063)} \approx 112$ feet per second squared during the time 0.063 second occupied in acquiring the gain. This gives -- 447 ft. per second per second average acceleration for portion $1..2$ of path, and since acceleration is constant throughout this portion, we have for portions of the ball's motion as it passes the screen $B$ is very nearly equal to this viz.-- 447 ft. per second per second (or "ft. per square second") By a similar process we find for portion $2..3$, viz.-- zero acceleration at screen $C$. Of course the reason why these results are merely approximate is that the spaces and times concerned, though relatively small, are not infinitesimal. [A recent English writer calls a unit of velocity a "speed," and a unit of acceleration a "hurry."] PARALLEL FORCES AND THE CENTRE OF GRAVITY. 1B CHAPTER II. PARALLEL FORCES AND THE CENTRE OF GRAVITY. 20. Preliminary Remarks.--Although by its title this section should be restricted to a treatment of the equilibrium of forces, certain propositions involving the composition and resolution of forces, without reference to the behavior of the body under their action, will be found necessary as preliminary to the principal object in view. As a rigid body possesses extension in three dimensions, to deal with a system of forces acting on it we require three co-ordinate axes: in other words, the system consists of "forces in space," and in general the forces are non-concurrent. In most problems in statics, however, the forces acting are in one plane; when this is so, they may be resolved into normal and parallel forces in a plane, of which the simplest case is that of two parallel forces. For the present the body on which the forces act will not be shown in the figure, but must be understood to be there (since we have no conception of forces independently of material bodies). The device will frequently be adopted of introducing into the given system two opposite and equal forces acting in the same line: evidently this will not alter the effect of the given system, as regards the rest or motion of the body. 21. Resultant of two Parallel Forces. CASE I.--The two forces have the same direction. Fig. 10. Let $P$ and $Q$ be the given forces, and $AB$ a line perpendicular to them ($P$ and $Q$ are supposed to have been transferred to the intersections $A$ and $B$). Put in at $A$ and $B$ two equal and opposite forces $S_1$ and $S_2$, combining them with $P$ and $Q$ to form $P'$. Fig. 10. 13 14 MECHANICS OF ENGINEERING. and $Q'$. Transfer $P$ and $Q$ to their intersection at $C$, and there resolve them again into $S$ and $P$, and $Q$. $S$ and $S$ annul each other at $C$; therefore $P$ and $Q$, acting along a straight line, $CD$, replace the two forces $P$ and $Q$ given, so that the resultant of the original two forces is a force $R = P + Q$, acting parallel to them through the point $D$, whose position must now be determined. The triangle $CAD$ is similar to the triangle shaded by lines, $\therefore P : S :: CD : z$; and $CDB$ is similar to the triangle shaded by dots, $\therefore S : Q :: a - x : CD$. Combining these, we have* $$\frac{P}{Q} = \frac{a-x}{x} \quad \text{and} \quad x = \frac{QS}{T+Q}$$ Now write this $$Rz = Qa,$$ and add $Re$, i.e., $Pe + Qe$, to each member, $e$ being the distance of $O$ (Fig. 10), any point in $AB$ produced, from $A$. This will give $(R + e)z = (P + Qe)(a - c)$, in which $c$, $a+c$, etc., are respectively the perpendiculars let fall from $O$ upon $P$, $Q$, and their resultant $R$. Any one of these products, such as $Pe$, is for convenience (since products of this form occur so frequently in Mechanics as a result of algebraic transformation) called the *moment* of the force about the arbitrary point $O$. Hence the resultant of two parallel forces of the same direction is equal to their sum, acts in their plane, in a line parallel to them, and at such a distance from any arbitrary point in that plane as may be determined by writing its moment about $O$ equal to that of the forces. The point of the two forces about $O$. $O$ is called a centre of moments, and each of the perpendiculars a *lever-arm*. CASE II.—Two parallel forces $P$ and $Q$ opposite directions. Fig. 11. By a process similar to the foregoing, we obtain $$R = P - Q$$ and $(P-Q)z = Qa$, i.e., $Rz = Qa$. Subtract each member of the last equation from either side, $(P-Q)z - Qa$, which gives $$R(c-x) = Pe - Q(a+c).$$ But $(c-x)$, $e$, and $(a+c)$ are respectively the perpendiculars, from --- That is, the resultant of two parallel forces pointing in the same direction divides the distance between them in inverse ratio of those forces. Fig. 11. PARALLEL FORCES AND THE CENTRE OF GRAVITY. 15 $O$, upon $R$, $P$, and $Q$. That is, $R(c-x)$ is the moment of $R$ about $O$; $Pc$, that of $P$ about $O$; and $Q(a+c)$, that of $Q$ about $O$. But the moment of $Q$ is subtracted from that of $P$, which corresponds with the fact that $Q$ in this figure would produce a rotation about $O$ opposite in direction to that of $P$. Having in view then, this imaginary rotation, we may define the moment of a force about a point as the indicated direction about the given point is against the hands of a watch; as negative when with the hands of a watch.* Hence, in general, the resultant of any two parallel forces is, in amount, equal to their algebraic sum, acts in a parallel direction in the same plane, while its moment, about any arbitrary point in the plane, is equal to the algebraic sum of the moments of the two forces about the same point. Corollary.--If each term in the preceding moment equations be multiplied by the secant of an angle (a, Fig. 13) thus: Diagram showing a parallelogram with vectors P, Q, R, S and angles α, β, γ. (using the notation of Fig. 13), we have $Pa \sec \alpha = P_a$, see $\alpha + P_a$, see $\alpha$, i.e., $Pb = Pb_1 + Pb_2$, in which $b_1$, $b_2$, and $\lambda$ are the oblique distances of the three lines of action from any point $O$ in their plane, and lie on the same straight line; $P$ is the resultant of the parallel forces $P_1$ and $P_2$. 29. Resultant of any System of Parallel Forces in Space.-- Let $P_1$, $P_2$, $P_3$, etc., be the forces of the system, and $x_1$, $y_1$, $z_1$, $x_2$, $y_2$, $z_2$, etc., the co-ordinates of their points of application as referred to an arbitrary set of three co-ordinate axes $X$, $Y$, and $Z$, perpendicular to each other. Each force is here re- * These two directions of rotation are often called counter clockwise, and clockwise, respectively. Fig. 13 Fig. 13 13 MECHANICS OF ENGINEERING. sticted to a definite point of application in its line of action (with reference to establishing more directly the fundamental equations for the co-ordinates of the centre of gravity of a body). The resultant $P'$ of any two of the forces, as $P_1$ and $P_2$, is $P' = P_1 + P_2$, and may be applied at the intersection of its line of action with the line $BD$, joining the points of application of $P_1$ and $P_2$, its components. Produce the latter line to $\mathcal{A}$, where it pierces the plane $XY$, and let $\delta_1$, $\delta_2$, and $\delta_3$ respectively, be the distances of $B$, $C$, $D$ from $\mathcal{A}$; then from the corollary of the last article we have $$P'v = P\delta_1 + P\delta_2;$$ but from stellar triangles $$b': b_1 : l_1 : x' : x_1 \therefore P'z' = P\delta_1 + P\delta_2.$$ Now combine $P'$, applied at $C$, with $P_2$ applied at $E$, calling their resultant $P''$ and its vertical co-ordinate $z''$, and we obtain $$P''z'' = P'z' + P\delta_2,$$ i.e., $P''z'' = P\delta_1 + P\delta_2 + P\delta_2,$ also $$z'' = P' + P = P + P_1 + P_2.$$ Proceeding thus until all the forces have been considered, we shall have finally, for the resultant of the whole system, $$R = P + P_1 + P_2 + \text{etc.};$$ and for the vertical co-ordinate of its point of application, which we may write $z'$, $$Rz = R\delta_1 + R\delta_2 + R\delta_3 + \text{etc.}$$ i.e., $$z' = \frac{R}{T}(\delta_1 + \delta_2 + \delta_3) = \frac{S(P)}{T}.$$ and similarly for the other co-ordinates. $$x' = \frac{S(P)}{T} \text{ and } y' = \frac{S(P)}{T}.$$ In these equations, in the general case, such products as $P\delta_n$, etc., cannot strictly be called moments. The point whose co- PARALLEL FORCES AND THE CENTRE OF GRAVITY. 17 ordinates are the $x$, $y$, and $z$ just obtained, is called the Centre of Parallel Forces, and its position is independent of the (common) direction of the forces concerned. **Example.—If the parallel forces are contained in one plane, and the axis $Y$ be assumed parallel to the direction of the forces, then each product like $P_x$ will be a moment, as defined in § 21; and it will be noticed in the accompanying numerical example, Fig. 14, that a detailed selection in the equation $$E_0 = P_x + P_y + \text{etc.}$$ having regard to the proper sign of each force and its distance, gives the result that if each product $P_x$ were first obtained numerically, and a sign affixed to the product considered as a moment about the point $O$. Let $P_x = -1$ lb.; $P_y = +3$ lbs.; $P_z = +3$ lbs.; $P_x = -6$ lbs.; $P_y = +1$ ft.; $z_x = +3$ ft.; $z_y = -2$ ft.; and $z_z = -1$ ft. Required the amount and position of the resultant $R$. In amount $R = 2P = -1 + 2 + 3 - 6 = -2$ lb.; i.e., it is a downward force of 2 lbs. As to its position, $$E_0 = 2(P_x)$$ gives $$(-2)x = (-1) \times (+1) + 2 \times 3 + (-2) \times (-6) = -1 + 6 + 12 - 12 = 5$$ Now from the figure by inspection, it is evident that the moment of $P_x$ about $O$ is negative (with the hands of a watch), and is numerically $-1$, i.e., its moment is negative; similarly, that of $P_y$ is seen to be positive, that of $P_z$ negative, that of $P_x$ positive; which agree with the results just found, that ($-9)x = -1 + 6 - 6 + 6 = -3$ ft. (Since a moment is a product of a force (lbs.) by a length (ft.), it may be called so many foot-pounds.) Next, solving for $x$, we obtain $$x = (-\frac{5}{2})$$ Thus, since $(+3)(-2) = -6$ lbs., the moment of the given forces is downward force of 9 lbs. acting in a line at 2.5 ft. to the left of the origin. Hence, body in question be a horizontal rod whose weight has been already included in the statement of forces, a support placed 2.5 ft. to the left of $O$ and capable of resisting at least 2 lbs. downward pressure will preserve equilibrium; and the pressure which it exerts... .18 MECHANICS OF ENGINEERING. against the rod must be an upward force, $P_e$ of 2 lbs., i.e. equal and opposite to the resultant of $F_p$, $F_q$, $F_r$, $F_s$. Fig. 15 shows the rod as a free body in equilibrium under the five forces. $P_e = -2 \text{ lbs.} = \text{the reaction of the support}$. Of course $P_e$ is one of a pair of equal and opposite forces, the other one is the pressure of the rod against the support, and would take its place among the forces acting on the support. **93. Centre of Gravity.** Among the forces acting on any rigid body at the surface of the earth is the so-called attraction of the latter (i.e., gravitation), as shown by a spring-balance, which indicates the weight of the body hung upon it. The weights of the different particles of any rigid body constitute a system of parallel forces (practically so, though actually slightly convergent). The point of application of the resultant of these forces is called the centre of gravity of the body, and may also be considered the centre of mass, the body being of very small dimensions compared with its distance from the centre. If $x$, $y$, and $z$ denote the co-ordinates of the centre of gravity of a body referred to three co-ordinate axes, the equations derived for them in § 22 are directly applicable, with slight changes in notation. Denote the weight of any particle of the body by $dG$, its volume by $dV$, by $\gamma$ its heaviness (rate of weight, see § 7) and its co-ordinates by $x$, $y$, and $z$; then, using the integral sign as indicating a summation of like terms for all the particles of the body, we have, for homogeneous bodies, $$\overline{x} = \frac{\int y dV}{\int dV}; \quad \overline{y} = \frac{\int x dV}{\int dV}; \quad \overline{z} = \frac{\int z dV}{\int dV}. \quad (1)$$ while, if the body is homogeneous, $\gamma$ is the same for all its elements, and being therefore placed outside the sign of summation, is cancelled out, leaving for homogeneous bodies ($V$ denoting the total volume) $$\overline{x} = \frac{\int x dV}{\int dV}; \quad \overline{y} = \frac{\int y dV}{\int dV}; \quad \overline{z} = \frac{\int z dV}{\int dV}. \quad (2)$$ Fig. 15. PARALLEL FORCES AND THE CENTRE OF GRAVITY. 19 Corollary.—It is also evident that if a homogeneous body is for convenience considered as made up of several finite parts, whose volumes are $V_1$, $V_2$, etc., and whose gravity co-ordinates are $x_1$, $y_1$, $z_1$; $x_2$, $y_2$, $z_2$; etc., we may write $$\bar{z} = \frac{V_1 + V_2 + \ldots}{V_1 + V_2 + \ldots} \quad \ldots \quad (3)$$ If the body is heterogeneous, put $G$, (weights), etc., instead of $V_1$, etc., in equation (3). If the body is infinitely thin homogeneous shell of uniform thickness = A, then $dV = k dF$ ($dF$ denoting an element, and $F$ the whole area of one surface) and equations (3) become, after cancellation, $$\bar{z} = \frac{f dF}{F}; \quad \bar{y} = \frac{f dy}{F}; \quad \bar{z} = \frac{f dz}{F} \quad \ldots \quad (4)$$ Similarly, for a homogeneous wire of constant small cross-section (i.e., a geometrical line, having weight), its length being $s$, and an element of length $ds$, we obtain $$\bar{w} = \frac{f ds}{s}; \quad \bar{y} = \frac{f dy}{s}; \quad \bar{z} = \frac{f dz}{s} \quad \ldots \quad (5)$$ It is often convenient to find the centre of gravity of a thin plate by experiment, balancing it on a needle-point; other shapes are not so easily dealt with. 24. Symmetry.—Considerations of symmetry of form often determine the centre of gravity of homogeneous solids without analysis, or limit it to a certain line or plane. Thus the centre of gravity of a sphere, or any regular polyhedron, is at its centre of figure; and the centre of gravity of the middle plane axis of a thin plate of the form of a circular polygon, in the centre of figure; or a straight wire of uniform cross-section, in the middle of its length. Again, if a homogeneous body is symmetrical about a plane, the centre of gravity must lie in that plane, called a plane of 20 MECHANICS OF ENGINEERING. gravity; if about a line, in that line called a line of gravity; if about a point, in that point. 25. By considering certain modes of subdivision of a homogeneous body, lines or planes of gravity are often made apparent. For example, a plate, the middle of the bases of a trapezoidal plate is a line of gravity, since it bisects all the strips of uniform width determined by drawing parallels to the bases; similarly, a line joining the apex of a triangular plate to the middle of the opposite side is a line of gravity. Other cases can easily be suggested by the student. 26. Problems. (1) Required the position of the centre of gravity of a fine homogeneous wire of the form of a circular arc, $AB$, Fig. 16. Take the origin at $O$ at the centre of the curve, and take $x$ and $y$ parallel to the axes. Let the length of the wire, $dx = 2\pi x$; $dy$ = element of arc. We need determine only the $x$, since evidently $y = 0$. Equations (3), § 23, are applicable here, i.e., $\bar{x} = \frac{\sqrt{r}dx}{s}$. From similar triangles we have $$ds : dy :: r : x ; \therefore ds = \frac{rdy}{x};$$ $$\therefore \bar{x} = \frac{r}{2a}\int_{-a}^{+a} \frac{2ra}{2a}, \text{i.e., chord } \times \text{radius } + \text{length of wire. For a semicircular wire, this reduces to } \bar{x} = 2r + w.$$ **Problem 2. Centre of gravity of trapezoidal (and triangular) thin plates, homogeneous, etc.—Prolonging the non-parallel sides of the trapezoid to intersect at $O$, which take as an origin, making the axis $X$ perpendicular to the bases $b_1$ and $b_2$. We may here use equations (4), § 23, and may take a vertical strip for our element of area, $dF$, in determining $\bar{x}$ for each point of such a strip has the same z. Now $dF = y + y'dz$, and** A diagram showing a wire with points labeled A, B, C, D, E, F. A B C D E F PARALLEL FORCES AND THE CENTRE OF GRAVITY. 21 from similar triangles $y+y'=\frac{b}{A}x$. Now $F_1=\frac{1}{2}(bh-bh_1)$, can be written $\frac{1}{2}\frac{b}{A}(h^*-h_1)$, and $\bar{x}=\frac{\int x'dF}{F'}$ becomes $$\left[\int_{h_1}^{h^*}x'dx\right]+\frac{1}{2}\frac{b}{A}(h^*-h_1)=\frac{2}{3}\frac{h^*-h_1}{h^*-A},$$ for the trapezoid. For a triangle $h_1=0$, and we have $\bar{x}=\frac{2}{3}h_1$; that is, the centre of gravity of a triangle is one third the altitude from the base. The centre of gravity is finally determined by knowing Fig. 17. Fig. 18. that a line joining the middles of $b$ and $b_1$ is a line of gravity; or joining $O$ and the middle of $b$ in the case of a triangle. **Problem 3. Sector of a circle. Thin plate, etc.—Let the notation, axes, etc., be as in Fig. 18. Angle of sector = $2\varphi$; $x=t$ Using polar co-ordinates, the element of area $dF$ (a small rectangle) = $pd\varphi \cdot dp$, and its $x=r\cos \varphi$; hence the total area = $$F=\int_{-\pi/2}^{\pi/2}\left[\int_{-\pi/2}^{\pi/2}pd\varphi\right]d\varphi=\int_{-\pi/2}^{\pi/2}\frac{r^2}{2}d\varphi=\frac{r^2}{2}\left[\frac{\pi}{2}-(-\frac{\pi}{2})\right];$$ i.e., $F=r^2a$. From equations (4), § 23, we have $$\bar{x}=\frac{1}{F}\int x'dF$$ $$=\frac{1}{F}\int\int r\cos \varphi pd\varphi d\varphi=\frac{1}{F}\int_{-\pi/2}^{\pi/2}\left[\cos \varphi \int p'd\varphi\right]d\varphi.$$ 22 MECHANICS OF ENGINEERING. (Note on double integration.--The quantity $$\left[\cos \varphi \int_0^{\pi} d\varphi\right]d\varphi,$$ is that portion of the summation $\int \int \cos \varphi d\varphi d\varphi$ which belongs to a single elementary sector (triangle), since all its elements (rectangles), from centre to circumference, have the same $\varphi$ and $d\varphi.$) That is, $$\frac{r}{r}\frac{1}{2}\int_{-\alpha}^{+\alpha}\cos \varphi d\varphi = \frac{r}{2}\sin \alpha = \frac{9}{3}\sin \alpha;$$ or, putting $\beta = 2\alpha = \text{total angle of sector}, x = \frac{4r}{3} = \frac{4r}{3}.$ For a semicircular plate this reduces to $x = \frac{4r}{3}$ (Note.--In numerical substitution the area $a$ and $\beta$ used above (unless sin or cos is prefixed) are understood to be expressed in circular measure ($x$-measure); e.g., for a quadrant, $\beta = \frac{\pi}{2} = 1.5707 + ;$ for $30^\circ, \beta = \frac{\pi}{6};$ or, in general, if $\beta$ in degrees $= 180^\circ$, then $\beta$ in $x$-measure $= \frac{\pi}{n}$.) Problem 4. Sector of a flat ring; thin plate, etc.--Treatment similar to that of Problem 3, the difference being that the limits of the interior integrations are $\int_0^{\pi}$ instead of $\int_{-\alpha}^{\alpha}$. Result, $$x = \frac{4}{3}\frac{r^2}{r^2 - r^2}\sin \frac{\beta}{2}.$$ PARALLEL FORCES AND THE CENTRE OF GRAVITY. 23 **Problem 5.** Segment of a circle; thin plate, etc.—Fig. 90. Since each rectangular element of any vertical strip has the same $x$, we may take the strip as $dF$ in finding $z$, and use $y$ as the half-height of the strip. $dF = 2ydz$, and from similar triangles $y : y' = (-dy) : dz$, i.e., $adz = -ady$. Hence from eq. (4), § 28, $$\bar{z} = \frac{fz dF}{F} = \frac{fz dy dz}{F} = \frac{-2y dy dz}{F} = \frac{9}{3P} \left[1 - y^2\right] = \frac{a^2}{3P}$$ but $a$ is the half-chord, hence, finally, $\bar{z} = \frac{(chord)^2}{12P}$. **Problem 6.** Trapezoid; thin plate, etc., by the method in the corollary of § 28; equations (3). Required the distance $z$ from the base $AB$. Join $DB$, thus dividing the trapezoid $ABCD$ into two triangles $ADB = F_1$, $b_1$, and $DBC = F_2$, whose gravity $x$'s are, respectively, $x_1 = \frac{h}{2}$ and $x_2 = \frac{h}{4}$. Also, $y_1 = \frac{h}{4}$, $y_2 = \frac{h}{4}$, and $F$ (area of trapezoid) = $\frac{h}{4}(A + B)$, Eq. (8) of § 25 gives $$Fz = F_1z + F_2z ;$$ hence, substituting $(b_1 + b_2)\bar{z} = \frac{h}{3}\left(b_1 + b_2\right)$$ The line joining the middles of $b_1$ and $b_2$ is a line of gravity, and is divided in such a ratio by the centre of gravity that the following construction for finding the latter holds good: Prolong each base to opposite distances, making them equal to the other base; join in two points thus found ; then draw lines with the other line of gravity at the centre of gravity of the trapezoid. Thus, Fig. 91, with $BE = b$, and $DF = b$, join $FE$, etc. A diagram showing a trapezoid ABCD with diagonals AC and BD intersecting at point O. The height h is shown along AB and CD. The area A is shown along AD and BC. The midpoints of AB and CD are marked as b1 and b2 respectively. The line joining these midpoints is shown as a dotted line. Fig. 90 A diagram showing a trapezoid ABCD with diagonals AC and BD intersecting at point O. The height h is shown along AB and CD. The area A is shown along AD and BC. The midpoints of AB and CD are marked as b1 and b2 respectively. The line joining these midpoints is shown as a dotted line. Fig. 91 24 MECHANICS OF ENGINEERING. PROBLEM 7. Homogeneous oblique cone or pyramid.— Take the origin at the vertex, and the axis $X$ perpendicular to the base (or base, if a frustum). In finding $x$ we may put $dV = volume$ of any lamina parallel to $YZ$, $F$ being the base of such a lamina, each point of the lamina having the same $x$. Hence, (equations (3), § 23), $$\bar{x} = \frac{1}{V} \int dV, \quad V = \int F dV = F d\bar{x};$$ but $$F : F_x :: x^* : h_x^*, \quad \therefore F = h_x^* x^*,$$ and $$V = \frac{F_1}{h_x^*} dx^* = \frac{F_1}{h_x^*} \int dx^* = \frac{F_1}{h_x^*} \int d\bar{x} = \frac{F_1}{h_x^*} \int L_1.$$ For a frustum, $\bar{x} = \frac{3}{4} h_x^* - h_x^*$; while for a pyramid, $h_x$, being $= 0$, $\bar{x} = \frac{3}{4} h_x^*$. Hence the centre of gravity of a pyramid is one fourth the altitude from the base. It also lies in the line joining the vertex to the centre of gravity of the base. Problem 8.—If the heaviness of the material of the above cone or pyramid varied directly as $x$, $y$, being its heaviness at the base $F_x$, we would use equations (1), § 23, putting $y = \frac{3}{4} x$; and finally, for the frustum, $$\bar{x} = \frac{4}{5} h_x^* - h_x^*,$$ and for a complete cone $\bar{x} = \frac{4}{5} h_x$. 27. The Centrobatic Method.—If an elementary area $dF$ be revolved about an axis in its plane, through an angle $\alpha < 2\pi$. Fig. 23. 25 PARALLEL FORCES AND THE CENTRE OF GRAVITY. 25 the distance from the axis being $=x$, the volume generated is $dV = ax dF$, and the total volume generated by all the $dF$'s of a finite plane figure whose plane contains the axis and which lies entirely on one side of the axis, will be $V = \int dF = ax dF$. But from § 38, $ax dF dx = ax dF$, as being the length of path described by the centre of gravity of the figure, we may write: The volume of a solid of revolution generated by a plane figure, lying on one side of the axis, equals the area of the figure multiplied by the length of curve described by the centre of gravity of the figure. A corresponding statement may be made for the surface generated by the revolution of a line. The arc $\alpha$ must be expressed in $x$ measure in numerical work. 27a. Centre of Gravity of any Quadrilateral—Fig. 25a. Construction: $ABCD$ being any quadrilateral. Draw the diagonals. On the long segment $DE$ of $DB$ lay off $DE = BK$, the shorter, to determine $E$; similarly, determine $N$ on the other diagonal, by making $AN = AK$. Bisect $EK$ in $H$ and $KN$ in $L$. Then $IL$ is a gravity-line, and $NL$ is the centre of gravity, C. Proof.—If being the middle of $DH$, and $AH$ and $HG$ having been joined, I the centre of gravity of the triangle $ABD$ is found on $AH$, by making $HI = \frac{1}{2}AH$; similarly, by making $HL = \frac{1}{2}HG$, L is the centre of gravity of triangle $BDG$. -.. IL is parallel to AO and is a gravity-line of the whole figure; and the centre of gravity C may be found on it if we can make $AL : CI :: arc ABD : arc BDO (\S 38)$. But since in triangles ABD and BDO (§ 38), AD and BD are proportional to the slant heights (equally inclined to DB) AK and KG, i.e., to GN and NA. Hence HN, which divides IL in the required ratio, contains C, so that C is a gravity-line. By similar reasoning, using the other diagonal, AG, and 26 MECHANICS OF ENGINEERING. the two triangles into which it divides the whole figure, we may prove $EM$ to be a gravity-line also. Hence the construction is proved. **Th. Example.—1. Required the volume of a sphere by the centrobate method.** A sphere may be generated by a semicircle revolving about its diameter through an arc $a = 2\pi$. The length of the path described by its centre of gravity is $\frac{4r}{3}\pi r^2$ (see Prob. 3, § 26), while the area of the semicircle is $\frac{4\pi r^2}{3}$. Hence by § 27, $$\text{Volume generated} = \frac{4r}{3}\pi \cdot \frac{4\pi r^2}{3} = \frac{4}{3}\pi r^3.$$ 2. Required the position of the centre of gravity of the sector of a flat ring in which $r_1 = 21$ feet, $r_2 = 30$ feet, and $\beta = 80^\circ$ (see Fig. 19, and § 26, Prob. 4). $$\sin \frac{\beta}{2} = \sin 40^\circ = 0.64279,$$ and $\beta$ in circular measure = $$\frac{80}{180}\pi = \frac{4}{9}\pi = 1.3962.$$ By using $r_1$ and $r_2$ in feet, $\bar{x}$ will be obtained in feet. $$\therefore \bar{x} = \frac{4}{3}\left(\frac{r_1^3 - r_2^3}{r_1^3 - r_2^3}\right) \cdot \sin \frac{\beta}{2} = \frac{4}{3} \cdot \frac{1961}{41} \cdot 0.64279 = 18.87 \text{ feet}.$$ STATICS OF A RIGID BODY. 27 CHAPTER III. STATICS OF A RIGID BODY. 28. Couples.—On account of the peculiar properties and utility of a system of two equal forces acting in parallel lines and in opposite directions, it is specially considered, and called a Couple. The arm of a couple is the perpendicular distance between the forces ; its moment, the product of this arm, by one of the forces. The axis of a couple is an imaginary line drawn perpendicular to its plane on that side from which the rotation appears positive (against the direction of the arrow). The ideal rotation mean, suggested by the position of the arrows; any actual rotation of the rigid body is a subject for future consideration.) In dealing with two or more couples the lengths of their axes are made proportional to their moments; in fact, by selecting a proper scale, numerically equal to these moments. E.g., in Fig. 24, the moments of the two couples there shown are $Pa$ and $Qb$; their axes $p$ and $q$ so laid off that $Pa : Qb :: p : q$, and that the ideal rotation may appear positive, viewed from the outer end of the axis. 29. No single force can balance a couple.—For suppose the couple $P$, $P$, could be balanced by a force $F$, then this, acting at some point $C$, ought to hold the couple in equilibrium. Draw $CO$ through $O$, the centre of symmetry of the couple, and make $OD = OC$. At $D$ put in two opposite and equal forces, $F_1$ and $F_2$, equal and parallel to $F$. The supposed equilibrium is indubitable. But if $F_1$, $F_2$, and A diagram showing a couple with arrows indicating rotation. The left arrow points upwards and to the right, while the right arrow points downwards and to the left. The text "Couple" is written below the diagram. Fig. 26 30 28 MECHANICS OF ENGINEERING. $P$ are in equilibrium, so ought (by symmetry about $O$) $S$, $P$, and $P'$ to be in equilibrium, and they may be removed without disturbing equilibrium. But we have left $T$ and $R'$, which are evidently not in equilibrium; -- the proposition is proved by this rectitude ad absurdum. Conversely a couple has no single resultant. 30. A couple may be transferred anywhere in its own plane. ---First, it may be turned through any angle $\alpha$, about any point of its arm, or of its arm produced. Let $(P', P)$ be a couple, $G$ any point of its arm (produced), and $\alpha$ any angle. Make $GC = GA$, $CD = AB$, and put in at $C$, $F$, and $P$, equal to $P$ (or $P'$), opposite to $G$, and perpendicular to $GC$; and at $D$, equal to $P$, similar to $D$. Now apply and combine $I'$ and $P'$ at $O$, $P'$ and $P$ at $O$; then evidently $R$ and $R'$ neutralize each other, leaving $P$, and $P'$ equivalent to the original couple $(P', P')$. The arm $CD = AB$. Secondly, if $G$ be at infinity, and $\alpha = 0$, the same proof applies, i.e., a couple may be moved parallel to itself in its own plane. Therefore, by a combination of the two transferrals, the proposition is established for any transferal in the plane. 31. A plane may be replaced by another of equal moment in a parallel plane.--Let $(P, P')$ be a couple. Let $CD$, in a parallel plane, be parallel to $AB$. At $D$ put in a pair of equal and opposite forces, $S_1$ and $S_2$, parallel to $P$ and each = $\frac{AE}{ED}P$. Similarly at $C$, $S_1$ and $S_2$ parallel to $P$ and each = $\frac{DE}{EC}P$. But, from similar triangles, $$\frac{AE}{ED} = \frac{BE}{EC} \therefore S_1 = S_2 = S_1 = S_2$$ STATICS OF A RIGID BODY. [Note.—The above values are so chosen that the intersection point $E$ may be the point of application of $(P + S_1)$, the resultant of $P$ and $S_1$; and also of $(P + S_2)$, the resultant of $P$ and $S_2$, as follows from § 31; thus (Fig. 38), $AE = P - S_1$, $DE = P - S_2$, and its moment about any centre of moments, as $E$, its own point of application, should equal the algebraic sum of the moments of its components about $E$; i.e., $R \times zero = P \cdot AE - S_1 \cdot DE - S_2 \cdot DE = P$.] Fig. 37. Replacing $P'$ and $S_1$ by $(P + S_1)$, and $P$ and $S_2$ by $(P + S_2)$, the latter results cancel each other at $E$, leaving the couple $(S_1, S_2)$ with an arm $CD$, equivalent to the original couple $P', P'$ with an arm $AB$. But, since $S_i = \frac{BE}{EC} \cdot P = \frac{AB}{CD} \cdot P$, we have $S_i \times CD = P \times AB$; that is, their moments are equal. 32. Transference and Transformation of Couples.—In view of the foregoing, we may state, in general, that a couple acting on a rigid body may be transferred to any position in any parallel plane, and may have the value of its forces and arms changed in any way so long as its moment is kept unchanged, and still have the same effect on the rigid body (as to rest or motion, not in distorting it). Corollaries.—A couple may be replaced by another in any position so long as their axes are equal and parallel and similarly situated with respect to their planes. A couple can be balanced only by another couple whose axis is equal and parallel to that of the first, and dissimilarly situated. For example, Fig. 39, $PA$ being $QB$, the rigid body $AB$ (here supposed without weight) is in equilibrium in each Fig. 38. 30 MECHANICS OF ENGINEERING. case shown. By "reduction of a couple to a certain arm $a$" is meant that for the original couple whose arm is $a'$, with forces each = $P'$, a new couple is substituted whose arm shall be $a$, and the value of whose forces $P$ and $P'$ must be computed from the condition $$Pa = Pa', \quad i.e., \quad P = Pa' + a.$$ Diagram showing two couples P and Q, with their respective axes p and q. Fig. 81 33. Composition of Couples.—Let $(P, P')$ and $(Q, Q')$ be two couples in different planes reduced to the same arm $AB = a$, which is a portion of the line of intersection of their planes. Thus in what follows, the original values of the component forces and arms of the two couples were they have been transferred and replaced in accordance with § 32, so that $P.AB$, the moment of the first couple, and the direction of its axis, $p$, have remained unchanged; similarly for the other couple. Combining $P$ with $Q$ and $P'$ with $Q'$, we have a resultant couple $(R, R')$ whose arm is also $AB$. The axes $p$ and $q$ of the component couples are proportional to $P.AB$ and $Q.AB$, i.e., to $P$ and $Q$, and contain the same angle as $P$ and $Q$. Therefore the parallelogram $p.q$ is a parallelogram $P.Q$: Ra. Also $r$ is evidently perpendicular to the plane of the resultant couple $(R, R')$, whose moment is Ra. Hence $r$, the diagonal of the parallelogram on $p$ and $q$, is the axis of the resultant couple. To combine two couples, therefore, we have only to combine their axes, as if they were forces, by a parallelogram, the diagonal being the axis of the resultant couple; the plane of this couple will be perpendicular to the Diagram showing two couples P and Q, with their respective axes p and q. Fig. 81 STATICS OF A RIGID BODY. 31 axis just found, and its moment bears the same relation to the moments of the component couples as the diagonal axis to the two component axes. Thus, if two couples, of moments $P_a$ and $Q_b$, lie in planes perpendicular to each other, their resultant couple will be represented by $v(P_a + Q_b)$. If these couples in different planes are to be combined, the axis of their resultant couple is the diagonal of the parallelo-piped formed on the axes, laid off to the same scale and pointing in the proper directions, the proper direction of an axis being away from the plane of its couple, on the side from which the couple appears of positive rotation. 34. If several couples lie in the same plane their axes are parallel and the axis of the resultant couple is their algebraic sum; and a similar relation holds for the moments; thus, in Fig. 24, the algebraic sum of all three couples is zero, i.e., $P_a + P_b + P_c = 0$, which shows us that a convenient way of combining couples, when all in one plane, is to call the moments positive or negative, according as the ideal rotations are against, or with, the hands of a watch, as seen from the same side of the plane; the sign of the algebraic sum will then show the ideal rotation of the resultant couple. 35. Composition of Non-concurrent Forces in a Plane.—Let $P_1$, $P_2$, etc., be the forces of the system; $x_1$, $y_1$, $z_1$, $y_2$, etc., the co-ordinates of their points of application; and $\alpha$, $\alpha_1$, ..., etc., their angles with the axis $X$. Replace $P_i$ by its components $X_i$ and $Y_i$, parallel to the arbitrary axis of reference. At the origin put in two forces, opposite to each other and equal and parallel to $X_i$; similarly for $Y_i$. (Of course $X_i = P_i \cos \alpha$ and $Y_i = P_i \sin \alpha$.) We now have $P_i$ replaced by two forces $X_i$, A diagram showing a force system with its components along X and Y axes. A diagram showing a force system with its components along X and Y axes. 32 MECHANICS OF ENGINEERING. and $X$, at the origin, and two couples, in the same plane, whose moments are respectively $-XY$, and $YX$, and are therefore (§ 34) equivalent to a single couple, in the same plane with a moment $(=YX - XY)$. Treating all the remaining forces in the same way, the whole system of forces is replaced by the force $2(X) = X + X_1 + \ldots$ at the origin, along the axis $X$; the force $2(Y) = Y + Y_1 + \ldots$ at the origin, along the axis $Y$; and the couple whose mom. $G = X(Y - XY)$, which may be called the couple $C$ (see Fig. 35), and may be placed anywhere in the plane. Now $2(X)$ and $2(Y)$ may be combined into a force $F$; i.e., $$R = (2X)^2 + (2Y)^2$$ and its direction-cosine is $\cos \alpha = \frac{2X}{R}$. Since, then, the whole system reduces to $C$ and $R$, we must have for equilibrium $R = 0$, and $G = 0$; i.e., for equilibrium $2X = 0$, $2Y = 0$, and $2(YX - XY) = 0$. . eq. (1) If $R$ alone is zero, the system reduces to a couple whose moment is $G = X(Y - XY)$; and if $G$ alone is zero the system reduces to a single force $R$, applied at the origin. If, in general, neither $R$ nor $G = 0$, the system is still equivalent to a single force, but not applied at the origin (as could hardly be expected, since the origin is arbitrary); as follows (see Fig. 35). Replace the couple $C$ by one of equal moment, $G$, with each force = $R$. Its arm will therefore be $\frac{G}{R}$. Move this couple in the plane so that one of its forces $R$ may cancel the $R$ already at the origin, thus leaving a single resultant $R'$ for the whole system, applied in a line at a perpendicular distance, $$c = \frac{G}{R},$$ from the origin, and making an angle $\alpha$ whose cosine = $\frac{2X}{R}$, with the axis $X$. 36. More convenient form for the equations of equilibrium of non-concurrent forces in a plane.—In (I.), Fig. 34, $O$ being STATICS OF A RIGID BODY. 33 any point and $a$ its perpendicular distance from a force $P$; put in at $O$ two equal and opposite forces $P$ and $P'$ and call to $P$, and we have $P$ replaced by an equal single force $P'$ at $O$, and a couple whose moment is $+Pa$. (II.) shows a similar construction, dealing with the $X$ and $Y$ components of $P$, so that in (II.) $P$ is replaced by single forces $X'$ and $Y'$ at $O$ Diagram showing a rigid body with forces P, P', X', Y' acting on it. Fig. 51. (and they are equivalent to a resultant $P'$, at $O$, as in (L), and two couples whose moments are +Ya and -Xy.) Hence, $O$ being the same point in both cases, the couple $Pa$ is equivalent to the two last mentioned, and, their axes being parallel, we must have $Pa = Ty - Xy$. Equations (1), § 35, for equilibrium, may now be written $$\begin{align*} 2X &= 0, \\ 2Y &= 0, \\ 2(Pa) &= 0. \end{align*}$$ (2) In problems involving the equilibrium of non-concurrent forces in a plane, we have three independent conditions, or equations, and these determine the unknown quantities. For practical solution, then, the rigid body having been made free (by conceiving the actions of all other bodies as represented by forces), and being in equilibrium (which it must be if at rest), we apply equations (2) literally; i.e., assuming an origin and two axes, equate the sum of the $X$ components of all the forces to zero; similarly for the $Y$ components; and then for the "moment-equation," having dropped a perpendicular from the origin upon each force, write the algebraic sum of the products obtained by multiplying each force by its perpendicular, or "lever-written" equal to zero, calling each product + or - according as the ideal rotation appears against, or with, the hands of a watch, as seen from the same side of the plane. (The converse convention would do as well.) Diagram showing a rigid body with forces P, P', X', Y' acting on it. Fig. 51. 34 MECHANICS OF ENGINEERING. Sometimes it is convenient to use three moment-equations, taking a new origin each time, and then the $\Sigma X = 0$ and $\Sigma Y = 0$ are superfluous, as they would not be independent equations. 37. Problems involving Non-concurrent Forces in a Plane—Remarks. The weight of a rigid body is a vertical force through its centre of gravity, downwards. If the surface of contact of two bodies is smooth the action (pressure, or force) of one on the other is perpendicular to the surface of contact. To cut off a portion of a body, and to pull out a pin, to make a body free, its tension must be inserted in the line of the cord, and in such a direction as to keep the small portion still fastened to the body. In case the pin of a hinge must be removed, to make the body free, its pressure against the ring being unknown in direction and amount, it is most convenient to represent it by its unknown components $X$ and $Y$, in known directions. In the following problems there is supposed to be no friction. If the line of action of an unknown (force) is horizontal, we shall assume it to be horizontal (backwards), assume a direction for it and adhere to it in all the three equations, and if the assumption is correct the value of the force, after elimination, will be positive; if incorrect, negative. Problem 1—Fig. 35. Given an oblique rigid rod, with two loads $G_1$ (its own weight) and $G_2$, required the reaction of the smooth vertical wall at $A$, and the direction and amount of the hinge-pressure at $O$. The reaction at $A$ must be horizontal ; call it $X'$. The pressure at $O$, which is unknown in direction, will have both its $X$ and $Y$ components un-known. The three unknowns, then, are $X$, $X'$, and $Y$, while $G_1$, $G_2$, $a_x$, $a_y$, and $h$ are known. The figure shows the rod as a free body, all the forces acting on it have been put in, and, since the rod is at rest, constitute a system of non-concurrent forces in a plane, ready for the conditions of equilibrium. Taking origin and axes as in the figure. A diagram showing a rigid rod with two loads \( G_1 \) and \( G_2 \). The rod is supported by a smooth vertical wall at point A. The reactions at A are denoted as \( X' \) (horizontal component) and \( Y \) (vertical component). The pressure at O is shown with components \( X \) and \( Y \). The diagram also includes labels for \( G_1 \), \( G_2 \), \( a_x \), \( a_y \), and \( h \). Pro. 35 STATICS OF A RIGID BODY. $$\mathbb{X} = 0 \text{ gives } X - X' = 0; \quad \mathbb{Y} = 0 \text{ gives } Y - G_1 - G_2 = 0; \quad \mathbb{Z}(\overline{PA}) = 0, \text{ about } O, \text{ gives } X + X' - G_3a_1 - G_4a_2 = 0. \text{(The moments of } X_a \text{ and } Y_a \text{ about } O \text{ are each zero.) By elimination we obtain } Y = G_1 + G_2; \quad X = X' = (G_3a_1 + G_4a_2) + k; \text{ while the pressure at } O = V X'_1 + Y'_1, \text{ and makes with the horizontal an angle whose tan } Y = Y'_1 + X'_1. $$ [N.R. A special solution for this problem consists in this, that the resultant of the three forces $G_1$, $G_2$, and $G_3$ intersects the line of $X'$ in a point which is easily found by § 3.] The hinge-pressure must pass through this point, since these three forces in equilibrium must be concurrent.] We might vary this problem by limiting $X'$ to a safe value, depending on the stability of the wall, and making $A$ an unknown. The three unknowns would then be $X_a$, $Y_a$, and $h$. **Problem 2—** Given two rods with loads, three hinges (or "pin-joints"), and all dimensions: required the three hinge-pressure; i.e., there are six unknowns, viz., three $X$ and three $Y$ components. We obtain three equations from each of the two free bodies in Fig. 37. The student may fill out the details. Notice the application of the principle of action and reaction at $B$ (see § 3). **Problem 3—** A Warren bridge-truss rests on the horizontal smooth abutment-surfaces in Fig. 38. It is composed of equal isosceles triangles; no piece is continuous beyond a joint, each of which is a pin connection. All loads are considered as acting at the joints so that each piece will be subjected to a simple tension or compression. Fig. 37. Fig. 38. Fig. 39. 36 MECHANICS OF ENGINEERING. First, required the reactions of the supports $V_1$ and $V_2$; these and the loads are called the external forces. $\Sigma(Pa)$ about $O = 0$ gives (the whole true is the free body) $$V_1 \cdot 3a + P_1 \cdot 4a - P_2 \cdot 4a = 0;$$ while $\Sigma(Pa)$ about $K = 0$ gives $$- V_1 \cdot 3a + P_1 \cdot 4a + P_2 \cdot 4a = 0;$$ ∴ $V_1 = 4(3P_1 + 3P_2 + P_2);$ and $V_2 = 4[P_1 + 3P_2 + 3P_2].$ Secondly, required the stress (thrust or pull, compression or tension) in each of the pieces $A$, $B$, and $C$ cut by the imaginary line $DE$. The stresses in the pieces are called internal forces. These appear in a system of forces acting on a free body only when a portion of the truss or frame is conceived separated from the remainder in such a way as to expose an internal plane of one or more pieces. Consider as a free body the portion on the left of $DE$ (that on the right would serve well), and let $A$, $B$, and $C$ would be found to act in directions opposite to those they have on the other portion ; see § 3). Fig. 39. The arrows (forces) $A$, $B$, and $C$ are not pointed yet. They, with $F_v$, $P_a$, and $P_w$ form a system in equilibrium. $\Sigma(Pa)$ about $O = 0$ gives $$\left(AA\right) - V_2a + P_1 \cdot 4a + P_2 \cdot 4a = 0.$$ Therefore the moment $(AD) = \frac{1}{4}(V_1 - 3P_1 - P_2)$, which is positive, since (from above) $\not F_v > 3P_1 + F_v$. Hence A must point to the left, i.e., is a thrust or compression, and is $$\frac{a}{2A}\left[4V_1 - 3P_1 - P_2\right].$$ Similarly, taking moments about $O_p$, the intersection of $A$ and $B$, we have an equation in which the only unknown is $C$, viz., $(CH) - V_1 \cdot 4a = 0$. ∴ $(CH) = \frac{a}{2}[3V_1 - 2P_1]$. A diagram showing a triangular truss with points labeled A, B, C, D, E, O, O_p, and arrows indicating forces. Fig. 39 STATICS OF A RIGID BODY. 37 a positive moment, since $3V_1$ is $>2P_1$; i.e., $C$ must point to the right, i.e., is a tension, and $\frac{d}{dx}(3V_1-2P_1)=0$. Finally, to obtain $B$, put $2(\text{vert. comp.)}=0$; i.e., $(B\cos\varphi+V_1-P_1-P_2=0$. But $\cos\varphi=\frac{P_1+P_2-V_1}{B}$; but (see foregoing value of $V_1$) we may write $$V_1=(P_1+P_2)-(4P_1+4P_2)+4P_v$$ $\therefore B\cos\varphi$ will be + (upward) or - (downward), and $B$ will be compression or tension, as $4P_1<\cos\varphi>(4P_2+4P_v)$, $$B=[(P_1+P_2-V_1)+\cos\varphi]=\sqrt{A^2+\frac{d^2}{dx^2}}[(P_1+P_2-V_1)]$$ **Problem 4.**—Given the weight $G_r$ of rod, the weight $G_a$, and all the geometrical elements (the student will assume a convenient notation); required the tension in the cord, and the amount and direction of pressure on hinge-pin. **Problem 5.**—Roof-truss; pin-connection; all loads at joints; wind-presures $W$ and $W_t$ normal to $OA$; required the three reactions or supporting forces (of the two horizontal surfaces and one vertical surface), and the stress in each piece. All geometrical elements are given; also $F_x$, $F_y$, $W$ (Fig. 40). 38. Composition of Non-concurrent Forces in Space.—Let $P_x$, $P_y$, etc., be the given forces, and $x_o$, $y_o$, $z_o$, $x_o$, $y_o$, $z_o$, etc., their points of application referred to an arbitrary origin and axes; $x_o$, $y_o$, $z_o$, etc., the angles made by their lines of application with $X$, $Y$, and $Z$. Fig. 40. Fig. 41. 38 MECHANICS OF ENGINEERING. Considering the first force $P$, replace it by its three components parallel to the axes, $X_1 = P \cos \alpha$, $Y_1 = P \cos \beta$, and $Z_1 = P \cos \gamma$ ($P$ itself is not shown in the figure). At $O$, and also at $A$, put a pair of equal and opposite forces, each equal and parallel to $Z_1$; $Z_1$ is now replaced by a single force $Z$, acting upward at the origin, and two couples, one in a plane parallel to $YZ$ and having a moment $= - Z_y$, (as we see it looking toward $O$ from a remote point on the axis + $X$), and another in a plane parallel to $XY$, having a moment $= + Z_x$ (seen from a remote point on the axis + $Y$). Similarly at $O$ and $C$ put in pairs of forces equal and parallel to $X_1$, and have $X_1$ at $B$, replaced by the single force $X$, at the origin, and the couples, one in a plane parallel to $XY$, and having a moment $= - X_y$, seen from a remote point on the axis + $Z$, the other in a plane parallel to $ZX$, and of a moment $= - X_z$, seen from a remote point on the axis + $Y$; and finally, by a similar device, $Y_1$ at $B$ is replaced by a force $Y$, at the origin, and the couples, parallel to the planes $XY$ and $YZ$, and having moments $\pm Y_x \pm Y_z$ respectively. (In Fig. 43 the single forces at the origin are broken lines, while the two forces constituting any one of the six couples may be recognized as being equal and parallel, of opposite directions, and both continuous, or both dotted.) We have, therefore, replaced the force $P$, by three forces $X_1$, $Y_1$, $Z_1$, and six couples (shown more clearly in Fig. 43); the couples have been transferred to symmetrical positions. Combining each two couples whose axes are parallel to $X$, $Y$, or $Z$ they can be reduced to three, viz., one with an $X$ axis and a moment = $Y_x - Z_y$; one with a $Y$ axis and a moment = $Z_x - X_z$; one with a $Z$ axis and a moment = $X_y - Y_x$. STATICS OF A RIGID BODY. Dealing with each of the other forces $P_0$, $P_1$, etc., in the same manner, the whole system may finally be replaced by three forces $\mathbf{ZX}$, $\mathbf{ZY}$, and $\mathbf{ZZ}$ at the origin and three couples whose moments are, respectively, $$L = 2(ZY - ZY) \text{ with its axis parallel to } X;$$ $$M = 2(ZZ - ZZ) \text{ with its axis parallel to } Y;$$ $$N = 2(XY - YZ) \text{ with its axis parallel to } Z.$$ The "axes" of these couples, being parallel to the respective co-ordinate axes $X$, $Y$, and $Z$, and proportional to the moments $L$, $M$, and $N$, respectively, the axis of their resultant $C$, whose moment is $G$, must be the diagonal of a parallelepiped constructed on the three component axes (proportional to) $L$, $M$, and $N$. Therefore, $G = \sqrt{L^2 + M^2 + N^2}$, while the resultant of $\mathbf{ZX}$, $\mathbf{ZY}$, and $\mathbf{ZZ}$ is $$R = \sqrt{(2XY)^2 + (2YZ)^2}$$ acting at the origin. If $\alpha$, $\beta$, and $\gamma$ are the direction-angles of $R$, we have $\cos \alpha = \frac{Z}{R}$, $\cos \beta = \frac{Y}{R}$, and $\cos \gamma = \frac{X}{R}$; while if $\lambda$, $\mu$, and $\nu$ are those of the axis of the couple $C$, we have $\cos \lambda = \frac{L}{G}$, $\cos \mu = \frac{M}{G}$, and $\cos \nu = \frac{N}{G}$. For equilibrium we have both $G = 0$ and $R = 0$; i.e., separately, six conditions, viz., $\mathbf{ZX} = 0$, $\mathbf{ZY} = 0$, $\mathbf{ZZ} = 0$; and $L = 0$, $M = 0$, $N = 0$. (1) Now, noting that $\mathbf{ZX} = 0$, $\mathbf{ZY} = 0$, and $\mathbf{ZZ} = 0$ are the conditions for equilibrium of the system of non-concurrent forces which would be formed by projecting each force of our actual system upon the plane XY, and similar relations for the planes YZ and ZZ, we may restate equations (1) in another form, viz., in practical problems, viz.: **Note:** If a system of non-concurrent forces is in equilibrium, the plane systems formed by projecting the given system upon each of three arbitrary co-ordinate planes will each be in equilibrium. But we can obtain only six independent 40 MECHANICS OF ENGINEERING. equations in any case, available for six unknowns. If $R$ alone = 0, we have the system equivalent to a couple $C$, whose moment = $G$; if $R$ alone = 0, the system has a single resultant $R$ applied at the origin. In general, neither $R$ nor $C$ being = 0, we cannot further combine $R$ and $C$ (as was done with non-concurrent forces in a plane) to produce a single resultant unless the axis of $C$ is parallel to the axis of $R$. Call that angle $\theta$. If, then, $\cos \theta = \cos \alpha \cos \lambda + \cos \beta \cos \mu + \cos y \cos v = 0 = \cos 90^\circ$, we may combine $R$ and $C$ to produce a single resultant for the whole system; acting in a plane containing $R$ and parallel to the plane of $C$ in a direction parallel to $R$, at a perpendicular distance $x = \frac{G}{R}$ from the origin and $= R$ in intensity. The condition that a system of forces in space have a single resultant is, therefore, substituting the previously derived values of the cosines, $(2X).L+(2Y).M+(2Z).N=0$. This includes the cases when $R$ is zero and when the system reduces to a couple. To return to the general case, $R$ and $C$ not being in the same plane, the composition of forces in space cannot be further simplified. Still we can give any value we please to $P$, one of the forces of the couple $C$, calculate the corresponding arm $a = J_0$, then transfer $C$ until one of the $P$s has the same point of application as $R$, and combine them by the parallelogram of forces. We thus have the whole system equivalent to two forces, viz., the second $P$, and the resultant of $R$ and the first $P$. These two forces are not in the same plane, and therefore cannot be replaced by a single resultant. 39. Problem. (Non-concurrent forces in space.)—Given all geometrical elements (including $\alpha$, $\beta$, $\gamma$, angles of $P$), also the weight of $Q$, and weight of apparatus $G$: A being a hinge whose pin is in the axis $Y$, O a ball-and-socket joint: required the amount of $P$ (lbs.) to preserve equilibrium, also the pressures STATICS OF A RIGID BODY. 41 (amount and direction) at $A$ and $O$; no friction. Replace $P$ by its $X$, $Y$, and $Z$ components. The pressure at $A$ will have A diagram showing a rigid body with forces applied to it. The body is shown in three different positions, labeled 1, 2, and 3. The forces are applied at points on the body, and the directions of the forces are indicated by arrows. The diagram also shows the moments about point $O$. The text below the diagram states that the six unknowns are $P$, $X$, $Y$, $Z$, $\alpha$, and $\beta$. The text also mentions that these six equations can be used to obtain the six unknowns. Fig. 44 $Z$ and $X$ components ; that at $O$, $X$, $Y$, and $Z$ components. The body is now free, and there are six unknowns. $\mathbf{XX}$, $\mathbf{YY}$, and $\mathbf{ZZ}$ give, respectively, $$P\cos\alpha + X_1 + X_2 = 0;$$ $$P\cos\beta + Y_1 = 0; \text{and} Z_1 + Z_2 - Q - G - P\cos y = 0.$$ As for moment-equations (see note in last paragraph), projecting the system upon $YZ$ and putting $\mathbf{Z}(Pa)$ about $O = 0$, we have $$-ZI + Qd + Ge + (P\cos y)b + (P\cos\beta)e = 0;$$ projecting it upon $ZX$, and putting $\mathbf{Z}(Pa)$ about $O = 0$, we have $$Qr - (P\cos e)a - (P\cos y)a = 0;$$ projecting on $XY$, moments about $O$ give $$X_1 + (P\cos y)b - (P\cos\beta)a = 0.$$ From these six equations we may obtain the six unknowns, $P$, $X$, $Y$, $Z$, $\alpha$, and $\beta$. If for any one of these a negative result is obtained, it shows that its direction in Fig. 44 should be reversed. 42 MECHANICS OF ENGINEERING. CHAPTER IV. STATICS OF FLEXIBLE CORDS. **4w Postulates and Principles.** The cords are perfectly flexible and inextensible. All problems will be restricted to one plane. Solutions of problems are based on three principles, viz: Prin. I.—The strain on a cord at any point can act only along the cord, or along the tangent if it be curved. Prin. II.—We may apply to flexible cords in equilibrium all the conditions for the equilibrium of rigid bodies; since, if the system of cords became rigid, it would still, with greater reason, be in equilibrium. Prin. III.—The conditions of equilibrium cannot be applied, of course, unless the system can be considered a free body, which is allowable only when we conceive to be put in, at the points of support or fastening, the reactions (upon the cord) of those points and the supports removed. These reactions having been put in, then consider the case in Fig. 45 in one plane. If we take any point, $p$, on the cord as a centre of moments, knowing that the resultant $R$ of the forces $P_a$, $P_b$ and $P_c$ situated on one side of $p$, must lie along the cord through $p$ (by Prin. 1), therefore we have $P_a p - P_b p = P_c p$, $R \times zero = 0$, and (equally well) $P_a p - P_b p - P_c p = 0$. That is, in a system of cords in equilibrium in a plane, if a centre of moments be taken on the cord, the algebraic sum of the mo- Fig. 45 46 STATICS OF FLEXIBLE CORDS. 43 ments of those forces situated on one side (either) of this point will equal zero. **41. The Pulley.**—A cord in equilibrium over a pulley whose axle is smooth has the same tension on both sides; for, Fig. 46, A diagram showing a pulley with a cord passing over it. The pulley is labeled "P" and the cord is labeled "C". The angle between the cord and the radius of the pulley is labeled "α". The tension in the cord is labeled "T". The force applied to the cord is labeled "F". The force applied to the pulley is labeled "P". The force applied to the block is labeled "B". The force applied to the ground is labeled "G". The force applied to the rope is labeled "R". The force applied to the rope is labeled "Q". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to the rope is labeled "O". The force applied to 44 MECHANICS OF ENGINEERING. **44. Weights Suspended by Fixed Knots.—Given all the geo- metrical elements in Fig. 50, and one weight, $G_1$; required the re- taining forces $H_a$, $H_b$, and the forces $H_2$, $V_2$, $V_3$, and $V_4$ at the points of support, so that equilibrium may obtain. $H_a$ and $V_a$ are the hori- zontal and vertical components of the tension in the cord at $O$; similarly $H_b$ and $V_b$ those at $n$. There are $n+2$ unknowns. From Prin. II we have $\lambda x = 0$, and $xT = v_1$ i.e., $H_a - H_n = 0$, and $\left[ G_1 + G_2 + \ldots\right] - \left[V_2 + V_3\right] = 0$. While from Prin. III, taking the successive knots, 1, 2, etc., as centres of moment, we have $$- V_{a1} + H_{b1} + G_1(x_a - x_b) = 0,$$ $$- V_{a2} + H_{b2} + G_1(x_a - x_c) = 0,$$ etc., for $n$ knots. Thus we have $n+2$ independent equations, a sufficient number, and they are all of the first degree (with reference to the unknowns), and easily solved. As a special solution, we may, by § 43, resolve $G_1$ in the directions of the first and sec- ond cord-segments, and obtain their tensions by a parallelogram of forces; then at the second knot, knowing the tension in the second segment, we may find that in the third and $G_2$ in like manner, and so on. Of course $H_a$ and $V_a$ are components of the tension in the first segment, $H_n$ and $V_n$ of that in the last. **45. The converse of the problem in § 44, viz., given the weights $G_1$, etc., $x$, and $y$, the lengths $a$, $b$, $c$, required $H_a$, $V_a$, $H_b$, $V_b$, and the coordinates $x_1$, $y_1$, $x_2$, $y_2$, etc. of the fixed knots when equilibrium exists, contain 2$n+2$ un- knowns. Station furnishes a $n+2$ equations (already given in § 44); while geometry gives the other n equations, one for each cord-segment, viz., $(x^1)^{'} + y^1 = a^1; (x^2 - x^1)^{'} + (y^2 - y^1)^{'} = b^1; \text{etc.}$ **STATICS OF FLEXIBLE CORDS.** However, most of these $2n+2$ equations are of the second degree; hence in the general case they cannot be solved. **46. Loaded Cord as Parabola.—If the weights are equal and infinitely small, and are intended to be uniformly spaced along the horizontal, when equilibrium obtains, the cord having no weight, it will form a parabola. Let $q = \text{weight per horizontal line}$. At $O$, let the curve in which the cord hangs, and $m$ any point. We may consider the portion $Om$ as a free body, if the reactions of the contiguous portions of the cord are put in, $H_x$ and $T$, and these (from Prin. I.) must act along the tangents to the curve at $O$ and $m$, respectively; i.e., $H_x$ is horizontal, and $T$ makes some angle $\varphi$ (whose tangent is $\frac{dy}{dx}$ etc.) with the axis $X$. Applying Prin. II, $$\Sigma X = 0 \text{ gives } T \cos \varphi - H_x = 0; \text{i.e., } \frac{d^2y}{dx^2} = H_x; \quad . . . \quad (1)$$ $$\Sigma Y = 0 \text{ gives } T \sin \varphi - qx = 0; \text{i.e., } \frac{dy}{dx} = qx; \quad . . . \quad (2)$$ Dividing (2) by (1), member by member, we have $\frac{dy}{dx} = \frac{qx}{H_x};$ :. $dy = \frac{q}{H_x} dx$, the differential equation of the curve; $$y = \int_{H_x}^{H_y} dx = \int_{H_x}^{H_y} \frac{dx}{\frac{q}{H_x}}; \text{or } x^2 = \frac{2H_y}{q}, y$$ the equation of a parabola whose vertex is at $O$ and axis vertical. Note.—The same result, $\frac{dy}{dx} = \frac{q}{H_x}$, may be obtained by considering that we have here (Prin. III.) a free rigid body acted on by three forces, $T$, $H_0$, and $R = qx$, acting vertically through the middle of the cord. The resultant force $R$ must be equal and opposite to $T$. Fig. 52. Evidently also the tangent-line bisects the axis $x$. 46 MECHANICS OF ENGINEERING. **47. Problem under § 48. [Case of a suspension-bridge in which the suspension-rods are vertical, the weight of roadway is uniform per horizontal foot, and later considered with that of the deck.] The deflection is very small, the load is generally furnished with a stiffening truss to avoid deformation under passing loads.] Given the span $=2b$, Fig. 53, $$y = \frac{v}{a} = g \text{ lbs. per horizontal foot; required the tension in the cable at } O_1 \text{ also at } m; \text{ and the length of cable needed. From the equation of the parabola } y = \frac{g}{2b}x^2, \text{ putting } x = b \text{ and } y = a, \text{ we have } H_1 = gb^2/2 \text{ so that } ZX = 0 \text{ gives } H_1 = H_2; \therefore \text{the tension at } M = \sqrt{H_1^2 + P^2} = \frac{1}{2}(gb^2 + F).$$ The semi-length, $Om$, of cable (from p. 88, Todhunter's Integral Calculus) is (letting $n$ denote $H_1 + 2b$, $b^2 + 4a$) $$Om = \sqrt{n^2 + a^2 + n}\log[(\sqrt{n^2 + n^2 + a^2} + \sqrt{n}]$$ **48. The Catenary.** A flexible, inextensible cord or chain, of uniform weight per unit of length, hung at two points, and supporting its own weight alone, forms a curve called the catenary. Let the tension $H_1$ at the lowest point or vertex be represented (for algebraic convenience) by the weight of an imaginary length, $c$, of similar cord weighing $q$ lbs. per unit of length, i.e., $H_1 = qc$; an actual portion of the cord, of length $s$, weighs $gs$ lbs. Fig. 54 shows as free and in equilibrium a portion of the curve of any catenary having its vertex at the origin. Required the equation of the curve. The load is uniformly spaced along the curve, and not horizontally, as in §§ 46 and 47. $$ZX = 0 \text{ gives } T_{2s}^{de} = gc.$$ Hence, by division, $ody = edx$, and equating, $cdy' = eddx'$ $$\therefore Y = 0 \text{ gives } T_{2s}^{dy'} = ge; \text{ while}$$ *For the "transformed catenary," see p. 395.* A diagram showing a catenary curve with a horizontal line representing \( ZX = 0 \), indicating no external force acting on the system. Fig. 54 **STATICS OF FLEXIBLE CORDS.** Put $dy' = de' - dx'$, and we have, after solving for $de$ $$de = \frac{ds}{\sqrt{x^2 + c^2}} \cdot ds = c \sqrt{\frac{ds}{x^2 + c^2}} = \int_0^y \log(x + \sqrt{x^2 + c^2}) \, ds,$$ (3) and $$e = x \cdot \log(x + \sqrt{x^2 + c^2}) + c, \quad . . .$$ a relation between the horizontal abscissa and length of curve. Again, in eq. (1) put $dx' = de' - dy'$, and solve for $dy'$. This gives $dy = \frac{ds}{\sqrt{x^2 + c^2}} = \frac{1}{2} \cdot \frac{d(x^2 + s)}{\sqrt{x^2 + c^2}}$. Therefore $$y = \int_0^s (\sqrt{x^2 + s})^{-1} d(x^2 + s),$$ (4) and finally $$y = \sqrt{x^2 + c} - c.$$ (5) Clearing of radicals and solving for $c$, we have $$c = (x^2 - y') + 9y.$$ (6) **Example.** A 40-foot chain weighs 240 lbs., and is so hung from two points at the same level that the deflection is 10 feet. Here, for $s = 20$ ft., $y = 10$; hence eq. (4) gives the parameter, $\epsilon = (400 - 100) / 20 = 15$ feet. $q = 240 / 40 = 6$ lbs. per foot; thus the tension in the middle is $T_h = qo = 6 \times 15 = 90$ lbs.; while the greatest tension is at either support and $\epsilon(90 + 150) = 150$ lbs. Knowing $c = 15$ feet, and putting $s = 20$ feet = half length of chain, we may compute the corresponding value of $x$ from eq. (3); this will be the half-span [log m = 2.30285] $\times$ (common log $m$)]. To derive $s$ in terms of $x$, transform eq. (2) in the sense in which $n = \log m$ may be transformed into $e^n$, clear of radicals, and solve for $s$, which gives $$s = \frac{e^{x^2} - e^{-x^2}}{e^{x^2} + e^{-x^2}},$$ (7) Again, eliminate $s$ from (3) by substitution from (5), transform as above, clear of radicals, and solve for $y + c$, whence $$y + c = \frac{e^{x^2} - e^{-x^2}}{e^{x^2} + e^{-x^2}},$$ (8) 48 MECHANICS OF ENGINEERING. which is the equation of a catenary with axes as in Fig. 54. If the horizontal axis be taken a distance $c$ below the ver- tex, then the ordinate is $y = x + c$, while $x$ remains the same; the last equation is simplified. If the span and length of chain are given, or if the span and deflection are given, $c$ can be determined from (5) or (6) only by successive assumptions and approximations. 4a. Addendum to § 53. Mass.—In Physics, the fundamental units are those of SPACE, involving a unit of length (and those of area and volume); TIME, involving a unit of time (and those of mass and force); MASS, involving a unit of mass, which (by Government decree) may be the quantity contained in one cubic centimeter of water at its freezing point; etc., a balance-beam being used to determine equal quantities of mass; while FORCE, involving a unit of force (and those of velocity and acceleration), involving the velocity of a moving mass, since it is proportional both to the mass and the acceleration; so that if we have a balance-beam to measure out a unit of mass to pro- duce unit acceleration in a unit of mass, so that we produce an acceleration $p = m$ in a mass $m$, then this balance-beam is itself a force, whose absolute value is an absolute unit. This called the dynamic measure of a force. Example: A man weighing 100 pounds will require about 1000 foot-pounds of energy to cause a mass of 400 grains to gain 30 velocity units in 2 sec.; i.e., $p = 100$ foot-pounds. The absolute value of this force is $\frac{1}{400} \times 100 = 2500$ alms. of force (or dynes, in C.G.S. system). In the fl-th-sec. system the absolute unit is called a poundal. In MECHANICS OF ENGINEERING, however, it is more convenient to regard the fundamental units as follows: SPACE, as ft., metre, etc., area and volume corresponding; TIME, as sec., minute, hour, day, etc., indicated by a gong-balance; Force, as lbs., grams, kilograms, tons, etc., indicated by a gong-balance; when for We assume a definite unit, a mode of measuring it being developed as follows. If we experiment (block on smooth table, for instance) we find that we can find a force $F$ acting on a body which will give it an acceleration $a$. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force of a body whose weight is the previous trial with a spring balance. We call this the "accelerating" motion in line of force PART II.—DYNAMICS. CHAPTER I. RECTILINEAR MOTION OF A MATERIAL POINT. 49. Uniform Motion implies that the moving point passes over equal distances in equal times; variable motion, that unequal distances are passed over in equal times. In uniform motion the distance passed over in a unit of time, as one second, is called the velocity (= v), which may also be obtained by dividing the length of any portion (= s) of the path by the time taken to take that portion (= t). The velocity is great; in variable motion, however, the velocity varies from point to point, its value at any point being expressed as the quotient of $ds$ (an infinitely small distance containing the given point) by $dt$ (the infinitely small portion of time in which $ds$ is described). 49a. By acceleration is meant the rate at which the velocity of a variable motion is changing at any point, and may be a uniform acceleration, in which case it equals the total change of velocity between any two points, however far apart, divided by the time of passage; or a variable acceleration, having a different value at every point, this value then being obtained by dividing the velocity-increment, $dv$, or gain of velocity in passing from one point to another, by the infinitely small time $dt$, the time occupied in acquiring the gain.* (Acceleration must be understood in an algebraic sense, a negative acceleration implying a decreasing velocity, or else that the velocity in a negative direction is increasing.) The foregoing applies to motion in a path or line of any form whatever, the distances mentioned being portions of the path, and therefore measured along the path. *See addendum on p. 12. 50 MECHANICS OF ENGINEERING. **50. Rectilinear Motion, or motion in a straight line.—The general relations of the quantities involved may be thus stated (see Fig. 55): Let $v$ = velocity of the body at any instant; $ds$ = distance traversed in a time $dt$; then $dv$ = gain of velocity at which $v$ is increasing at any instant. All these may be variable; and $t$ is taken as the independent variable, i.e., time is conceived to change by equal small increments, each = $dt$; hence two consecutive $ds$ will not in general be equal, their difference being called $d^{\prime}ds$. Evidently $d^{\prime}ds$ = zero, i.e., $dt$ is constant. Since $\frac{1}{dt}$ = number of instants in one second, the velocity at any instant (i.e., the distance which would be described at that rate in one second) is $v = ds \cdot \frac{1}{dt} ; \therefore v = \frac{ds}{dt}$ . . . . . . . . . . . . . . . . . . (I.) Similarly, $p = dv \cdot \frac{1}{dt}$ and (since $dv = d\left(\frac{ds}{dt}\right) = \frac{d^{\prime}ds}{dt}$) $\therefore p = \frac{dv}{dt} = \frac{d^{\prime}p}{dt}$ . . . . . . . . . . . . . . . . . (II.) Eliminating $dv$, we have also $vdv = pdp$. . . . . . . . . . . (III.) These are the fundamental differential formulae of rectilinear motion (for curvilinear motion we have these and some in addition) as far as kinematics, i.e., as far as space and time, is concerned. The consideration of the mass of the material point and the forces acting upon it will give still another relation (see § 85). **51. Rectilinear Motion due to Gravity.—If a material point fall freely in vacuo, no initial direction other than vertical having been given to its motion, many experiments have** 50 RECTILINEAR MOTION OF A MATERIAL POINT. 51 shown that this is a uniformly accelerated rectilinear motion in a vertical line having an acceleration (called the acceleration of gravity) equal to 32.2 feet per square second, or 9.81 metres per square second; i.e., the velocity increases at this constant rate in a downward direction, or decreases in an upward direction. [Note.--By "square second" it is meant to lay stress on the fact that an acceleration (being $dv/dt$) is in quality equal to one dimension of length divided by two dimensions of time. E.g., if instead of using the foot and second, we use the inch and minute, then the value of $g$ will = 20.3 × 6000; whereas a velocity of say six feet per second would = $6 \times 60$ feet per minute. The value of $g$ = 32.2 implies the units foot and second, and is sufficiently exact for practical purposes.] 52. Free Fall in Vacuum--Fig. 56. Let the body start at O with an initial downward velocity $=v$. The acceleration is constant and = $+g$. Reckoning both time and distance (+ downwards) from $O$, required the values of the variables $e$ and $v$ after any time $t$. From eq. (II), § 50, we have $+g = dv/dt$; i.e., $do = gdt$, in which the two variables are separated. Hence $\int_0^t do = g\int_0^t dt$; i.e., $v = g\int_0^t t; \text{or } v - c = gt - 0$; and finally, $v = c + gt$ (Notice the correspondence of the limits in the foregoing operation; when $t = 0$, $v = +c$.) From eq. (I.), § 50, $v = ds/dt$; substituting from (1), $ds = (c + gt)dt$, in which the two variables $v$ and $t$ are separated. $\therefore \int_0^t ds = c\int_0^t dt + g\int_0^t dt$, i.e., $\frac{ds}{dt} = c + g\frac{dt}{dt}$, or $s = ct + \frac{gt^2}{2}$ Again, eq. (III.), § 50, $vdt = gdt$, in which the variables $v$ and $t$ are already separated. $\therefore \int_0^t vds = g\int_0^t ds$; or $\frac{ds}{dt} = g\frac{dt}{dt}$; i.e., $(v^2 - c^2) = gs$, or $s = \frac{v^2 - c^2}{2g}$ (3) A diagram showing a free fall in vacuum. 53 MECHANICS OF ENGINEERING. If the initial velocity = zero, i.e., if the body falls from rest, eq. (3) gives $s = \frac{v^2}{2g}$ and $v = \sqrt{\frac{2gh}{g}}$. [From the frequent re- enrence of these forms, especially in hydraulics, $\frac{v^2}{2g}$ is called the "height due to the velocity" $v$; i.e., the vertical height through which the body must fall from rest to acquire the velocity $v$; while, conversely, $\sqrt{\frac{2gh}{g}}$ is called the velocity due to the height or head A.] By eliminating $g$ between (1) and (3), we may derive another formula between three variables, $s$, $v$, and $t$, viz., $$s = \left( \frac{v^2}{2g} + vt \right).$$ (4) 53. Upward Throw.—If the initial velocity were in an up- ward direction in Fig. 56 we might call it $+c$, and introduce it with a negative sign in equations (1) to (4), just derived; but for variety let us call the upward direction $+c$, in which case an upward initial velocity would be $-c$, while the acceleration = $-g$, constant, as before. (The motion is supposed confined within such a small range that $g$ does not sensibly vary.) Fig. $$\begin{array}{l} +5 \\ 57. \text{ From } p = dv/dt \text{ we have } dv = -gdz \text{ and} \\ \int_{0}^{h} dv = -g\int_{0}^{h} dt; \therefore -c = -gt; \text{or } v = c - gt. (1)a \\ \int_{0}^{h} dt = c\int_{0}^{h} dt; \text{or } v = ct - gdt. (2)a \\ \therefore vdo = pds \text{ gives } \int_{0}^{h} vds = -g\int_{0}^{h} ds, \text{ whence} \\ \frac{1}{2}(v^2 - c^2) = -gs, \text{ or finally, } s = \frac{c^2 - v^2}{2g}. (3)a \\ \end{array}$$ And by eliminating $g$ from (1a) and (3)a, $$s = \left( \frac{v^2}{2g} + vt \right).$$ (4)a The following is now easily verified from these equations : the body passes the origin again ($e=0$) with a velocity $= -c$, after a lapse of time $= 2c/g$. The body comes to rest for *In Hydraulics A is used instead of $s$.* A diagram showing a body falling under gravity. RECTILINEAR MOTION OF A MATERIAL POINT. 53 at instant) (put $v = 0$) after a time $= c + g$, and at a distance $s = c^2 + 2g$ (height due to velocity $c$) from $O$. For $t > c + g$, $v$ is negative, showing a downward motion ; for $t > 2c + g$, $s$ is negative, i.e., the body is below the starting-point, while the rate of change of $v$ is constant and $= -g$ at all points. 54. Newton's Laws.—As showing the relations existing in general between the motion of a material point and the actions (forces) of other bodies upon it, experience furnishes the following three laws or statements as a basis for dynamics : (1) A material point under no forces, or under balanced forces, remains in a state of rest or of uniform motion in a right line. (This property is often called Inertia.) (2) If the forces acting on a material point are unbalanced, an acceleration of motion is produced proportional to the resultant force and in its direction. (3) Every action (force) of one body on another is always accompanied by an equal, opposite, and simultaneous reaction. (This was interpreted in § 3.) As all bodies are made up of material points, the results obtained in Dynamics of a Material Point serve as a basis for the Dynamics of a Rigid Body, of Liquids, and of Gases. 55. Mass.—If a body is to continue moving in a right line, the resultant force $P$ at all instants must be directed along that line (otherwise it would have a component deflecting the body from its straight course). In accordance with Newton's second law, denoting by $p$ the acceleration produced by the resultant force ($g$ being the body's weight), we must have the proportion $P : g :: p : g$; i.e., $$P = \frac{G}{g} p \quad \ldots \quad \text{or} \quad P = Mp. \quad \ldots \quad (\text{IV.})$$ Eq. IV. and (I.), (II.), (III.) of § 50 are the fundamental equations of Dynamics. Since the quotient $G : g$ is invariant. * See Addendum on p. 48. A page from a book discussing Newton's laws of motion. ble, wherever the body be moved on the earth's surface ($G$ and $g$ changing in the same ratio), it will be used as the measure of the mass $M$ or quantity of matter in the body. In this way it will frequently happen that the quantities $G$ and $g$ will appear in problems where the weight of the body, i.e., the force of the earth's attraction upon it, and the acceleration due to gravity have no direct connection with the circumstances. No name will be given to the unit of mass, it being always understood that the fraction $G/g$ will be put for $M$ before any numerical substitution is made. From (IV.) we have, in words, accelerating force = mass x acceleration; also, acceleration = accelerating force - mass. 58. Uniformly Accelerated Motion—If the resultant force is constant as time elapses, the acceleration must be constant (from eq. (IV.), since of course $M$ is constant) and $P = M$. The motion therefore will be uniformly accelerated, and we have only to substitute + $p$ (constant) for $g$ in eqs. (1) to (4) of § 52 for the equations of this motion, the initial velocity being = $e$ (in the line of the force). $$v = e + pt \quad . . . \quad (1); \quad s = et + \frac{1}{2}pt^2; \quad . . . \quad (2)$$ $$e = \frac{(v - e)}{2p}; \quad . . . \quad (3), \quad \text{and} \quad s = \frac{1}{2}(c + v)t \quad . . . \quad (4)$$ If the force is in a negative direction, the acceleration will be negative, and may be called a retardation; the initial velocity should be made negative if its direction requires it. 57. Examples of Unif. Acc. Motion—Example 1. Fig. 58. A small block whose weight is 4 lb. has already described a distance $AO$ = 45 inches over a smooth portion of a horizontal table in two seconds; at $O$ it encounters a rough portion, and a consequent constant friction of 2 oz. Required the distance described beyond $O$, and the time occupied in coming to rest. Since we shall use $\frac{3}{2}$ for $g$, times must be in seconds, and distances in feet; as to the unit A diagram showing a block moving horizontally on a table. The block starts from point A and moves to point O in 2 seconds. At point O, it encounters a rough portion and comes to rest after traveling a certain distance beyond O. Fig. 58 RECTILINEAR MOTION OF A MATERIAL POINT. 58 A force, as that is still arbitrary, say ounces. Since $AO$ was smooth, it must have been described with a uniform motion (the resistance of the air being neglected); hence with a velocity = 4 ft. + 2 sec. = 2 ft. per sec. The initial velocity for the retarded motion, then, is $v = +2$ at $O$. At any point beyond $O$ the acceleration = force + mass = (-2 oz.) + (8 oz. + 32 ft./sec.²) = 6.05 ft./sec. squared, i.e., $p = -8.05$ constant; hence the retardation is uniform (retarded here), and we may use the formulae of § 56 with $a = +2$, $p = -8.05$. At the end of the motion $v$ must be zero, and the corresponding values of $s$ and $t$ may be found by putting $v = 0$ in equations (3) and (1), and solving for $s$ and $t$ respectively: thus from (3), $t = \frac{1}{4}(-4) + (-8.05)$, i.e., $t = 0.248 + t$, which must be feet; while from (1), $s = (-2) + (-8.05) = 0.248 + s$, which must be seconds. Example (Algebraic).—Fig. 59. The two masses $M_1 = G_1 + g$, and $M_2 = G_2 + g$ are connected by a flexible, inextensible cord. Table smooth. Required the acceleration common to the two rectilinear motions, and the tension in the string $S$, Fig. 59. there being no friction under $G_1$, none at the pulley, and no mass in the latter or in the cord. At any instant of the motion consider $G_1$ free (Fig. 60), $N$ being the pressure of the table against $G_1$. Since the motion is in a straight right line ($x$-rect., compone) = 0, i.e., $N - G_1 = S$, which determines $N$, and the only horizontal force (and resultant of all the forces) = $M_1 p$, i.e., $$S = G_1 p + \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \\ S = G_1 p + N - G_1 = S\end{equation*} At the same instant of the motion consider $G_2$ free (Fig. 61); the tension in the cord is the same value as above = $S$. The accelerating force is $G - S$, and $$\therefore = M_2 p,\text{ or } G - S = (G + g)p.\end{equation*} (9) 56 MECHANICS OF ENGINEERING. $$\frac{v^2}{g} = \frac{(Gg) + (G + G_0)}{g} = \text{constant}; \text{hence each motion is uniformly accelerated, and we may employ equations (1) to (4) of § 56 to find the velocity and distance from the starting-point, at the end of any assigned time } t, \text{ or vice versa. The initial velocity must be known, and may be zero.}$$ Also, from (1) and (2) of this article, $$S = (GG_0) + (G + G_0) = \text{constant}.$$ Example 3.—A body of 22 short tons weight is acted on during $\frac{1}{4}$ minute by a constant force $P$. It had previously described 3164 yards in 180 seconds under no force; and subsequently, under no force, describes 9900 inches in $\frac{1}{4}$ of an hour. Required the value of $P$. Ans. $P = 22.1$ lbs. Example 4.—A mass of 1 ton weight, having an initial velocity of 48 inches per second; is acted on for $\frac{1}{4}$ minute by a force of 400 avoirdupois ounces. Required the final velocity. Ans. 10.037 ft. per sec. Example 5.—Initial velocity, 80 feet per second; mass weighs 0.30 of a ton. A resistance of 112$\frac{1}{2}$ lbs. retards it for $\frac{1}{4}$ of a minute. Required the distance passed over during this time. Ans. 256.5 feet. Example 6.—Required the time in which a force of 600 avoirdupois ounces will increase the velocity of a mass weighing 14 tons from 480 feet per minute to 340 inches per second. Ans. 30 seconds. Example 7.—What distance is passed over by a mass of (0.6) tone weight during the overcoming of a constant resistance (friction), if its velocity, initially 144 inches per sec., is reduced to zero in 8 seconds. Required, also, the friction. Ans. 48 ft. and 55 lbs. Example 8.—Before the action of a force (value required) a body of 1 ton had described uniformly 950 ft. in 12 minutes. Afterwards it describes 1650 feet uniformly in 180 seconds. The force acts 30 seconds. $P = ?$ Ans. $P = 178$ lbs. A diagram showing a body moving with uniform acceleration. RECTILINEAR MOTION OF A MATERIAL POINT. 57 58. Graphic Representations. Unif. Acc. Motion.—With the initial velocity = 0, the equations of § 56 become $$v = pt, \quad (1)$$ $$s = vt^2 + \frac{1}{2}pt^2, \quad (2)$$ $$s = v^2 + 2pt, \quad (3)$$ and $$s = \frac{1}{2}vt^2 + pt^2, \quad (4)$$ Eqs. (1), (2), and (3) contain each two variables, which may graphically be laid off to scale as co-ordinates and thus give a curve corresponding to the equation. Thus, Fig. 62, in (I), we Fig. 62. have a right line representing eq. (L); in (II), a parabola with axis parallel to $s$, and vertex at the origin for eq. (2); also a parabola similarly situated for eq. (3). Eq. (4) contains three variables, $s$, $v$, and $t$. This relation can be shown in (I), $s$ being represented by the ordinate, but this alone does not show that (II) and (III) have this advantage, that the axis $OS$ may be made the actual path of the body. [Let the student determine how the origin shall be moved in each case to meet the supposition of an initial velocity = $+c$ or $-c$.] 59. Variably Accelerated Motions.—We here restate the equations $$v = \frac{ds}{dt}, \quad (L); p = \frac{de}{dt} = \frac{ds}{dt}, \quad (IL); vdp = pds, \quad (III);$$ and resultant force $$P = Mp, \quad (IV);$$ which are the only ones for general use in rectilinear motion. PROBLEM 1.—In pulling a mass $M$ along a smooth, horizontal table, by a horizontal cord, the tension is so varied that $s = At^2$ (not a homogeneous equation; the units are, say, the foot and second). Required by what law the tension varies. 58 MECHANICS OF ENGINEERING. From (L) $v = \frac{ds}{dt} = \frac{d(4s)}{dt} = 12t$; from (II), $p = \frac{d(12t)}{dt} = 24t$; and (IV). The tension $P = Mp = 24Mt$, i.e., varies directly as the time. Problem 2. "Harmonic Motion," Fig. 63.--A small block Diagram showing a small block attached to two horizontal elastic cords, with a spring in between. on a smooth horizontal table is attached to two horizontal elastic cords (and they to pegs) in such a way that when the block is at $O$, each cord is straight but not tense; in any other position, as $m$, one cord is tense, the other slack. The cords are alike in every respect, and, as with springs, the tension varies directly with the elongation ($= s$ in figure). If for an elongation $s$, the tension is $T_s$, then for any elongation $s$ it is $T = T_s + s$. The acceleration of the point $m$, then, is $a = \frac{ds}{dt} = M\frac{dT_s}{dt} - Ms$. For $T_s = Ma$, which we put $p = -as$, $a$ being a constant. Required the equations of motion, the initial velocity being $+c$, at $O$. From eq. (III) $$vds = -ads; \therefore \int_0^s vds = -a\int_0^s ads,$$ i.e., $\frac{d(v^2 - c^2)}{dt} = -4as^2; \text{ or, } v^2 = c^2 - 4as^2.$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) From (L), $dt = ds + v$; hence from (1), $$\int_0^s dt = \int_0^s (ds + v^2 - 4as^2),$$ or $$t = \frac{1}{\sqrt{a}}\int_0^s \frac{d(v^2 - 4as^2 + c)}{\sqrt{1 - (v^2 + c)^2}} = \frac{1}{\sqrt{a}}\sin^{-1}\left(\frac{v\sqrt{a}}{\sqrt{c}}\right)$$ $$= \frac{1}{\sqrt{a}}\sin^{-1}\left(\frac{\sqrt{a}}{\sqrt{c}}\right).$$ . . . . RECTILINEAR MOTION OF A MATERIAL POINT. 59 Inverting (3), we have $s = (c + \sqrt{a}) \sin(t + \theta)$, . . . . (3) Again, by differentiating (8), see (1), $w = c \cos(t + \theta)$ (4) Differentiating (4), see (II.), $p = -c \sqrt{a} \sin(t + \theta)$. . . . (5) These are the relations required, but the peculiar property of the motion is made apparent by inquiring the time of passing from $O$ to a state of rest; i.e., put $v = 0$ in equation (4), we obtain $t = \pi + \sqrt{a}$, or $\frac{\pi}{2} + \sqrt{a}$, or $\frac{\pi}{4} + \sqrt{a}$, and so on, while the corresponding values of $s$ (from equation (3)), are $+ (\frac{\pi}{2} + \sqrt{a}) - (\frac{\pi}{2} + \sqrt{a}) + (\frac{\pi}{2} + \sqrt{a})$, and so on. This shows that the body vibrates equally on both sides of $O$ in a cycle or period whose duration is $2\pi + \sqrt{a}$, and is independent of the initial velocity given it at $O$. Each time it passes $O$ the velocity is zero, and the time since the start is $\pi w + \sqrt{a}$, in which $w$ is any whole number. At the extreme point $p = \pm c \sqrt{a}$, from eq. (5). If then a different amplitude be given to the oscillation by changing $c$, the duration of the period is still the same, i.e., the vibration is *isochronal*. The motion of an ordinary pendulum is nearly that of a cycloidal pendulum exactly, harmonic. If the crank-pin of a reciprocating engine moved uniformly in its circular path, the piston would have a harmonic motion if the connecting-rod were infinitely long, or if the design in Fig. 64 were used. (Let the student prove this from eq. (3).) Let $2r$ = length of stroke, and $c$ = the uniform velocity of the crank-pin, and $M$ = mass of the piston and rod $AB$. Then the velocity of $M$ at mid-stroke must be $c$, at the dead-points, zero; its acceleration at mid-stroke zero; at the dead-points the acc. = $\sqrt{a}$, and $d^2 = r = c + \sqrt{a}$ (from eq. (3)). $\therefore d^2 = c + r$, and the acc. at a dead point (the maximum acc.) A diagram showing a crank mechanism with a piston moving up and down in a cylinder. Fig. 64 60 MECHANICS OF ENGINEERING. $$=e+r.$$ Hence on account of the acceleration (or retardation) of $M$ in the neighborhood of a dead-point a pressure will be exerted on the crank-pin, equal to mass $x$ acc. = $Mc + r$ at these points, independently of the force transmitted due to steam-pressure on the piston-head, and makes the resultant pressure on the pin at $C$ smaller, and at $D$ larger than it would be if the "inertia" of the piston and rod were not thus taken into account. We shall consider this case by the method, considering $AB$ free immediately after passing the dead-point. Fig. 65. The forces acting are: $G$, the weight; $N$, the pressures of the guides; $P$, the known effective steam-pressure on piston-head; and $P'$, the unknown pressure of crank-pin on side of point. There is no change of motion vertically; i.e., $a + N - G = 0$, and the resultant force is $P - P' = \text{mass} x$ acc. = $Mc + r$. Similarly at other dead-point we would obtain $P' = P + Mc + r$. In high-speed engines with heavy pistons, etc., $Mc + r$ is no small item. [The upper half-revol., alone, is here considered.] Problem 3.--Supposing the earth at rest and the resistance of the air to be null, a body is given an initial upward vertical velocity of $c$. Required the velocity at any distance $s$ from the centre of the earth, whose attraction varies inversely as the square of the distance. Fig. 66. See Fig. 66.--The attraction on the body at the surface of the earth where $s = r$, the radius, is its weight $G$; at any point $m$ it will be $G(r^2 + s^2)$, while its mass is $G + g$. Hence the acceleration at $m = p = (-P) + M$ $$= -g(r^2 + s^2).$$ Take equation III, $\frac{dv}{ds} = pds,$ and we have $$\frac{dv}{ds} = -gr^2\int_0^s ds; ;$$ $$\int_0^s dv = -gr^2\int_0^s ds; ; \frac{1}{r}\int_0^s dv = -gr^2\int_0^s \frac{1}{r};$$ i.e., $\frac{1}{r}(v^2 - c^2) = -gr^2(\frac{1}{r} - \frac{1}{r}); \ldots \quad (1)$ RECTILINEAR MOTION OF A MATERIAL POINT. 61 Evidently $v$ decreases, as it should. Now inquire how small a value $c$ may have that the body shall never return ; i.e., that $v$ shall not = 0 until $s = \infty$. Put $c = 0$ and $s = \infty$ in (1) and solve for $c$; and we have $$c = \sqrt{2gr} = \sqrt{g \times 32.2 \times 2100000},$$ = about 3680 ft. per sec. or nearly 7 miles per sec. Conversely, if a body be allowed to fall from rest, toward the earth, the velocity with which it would strike the surface would be less than 7 miles per second through whatever distance it may have fallen. If a body were allowed to fall through a straight opening in the earth passing through the centre, the motion would be harmonic, since the attraction and consequent acceleration now vary directly with the distance from the centre. See Prob. 2. This supposes the earth homogeneous. PROBLEM 4.--Steam working expansively and raising a weight. Fig. 67. The steam works without friction in a vertical cylinder. Let $S =$ total steam-pressure on the underside of the piston; the weight $G$, of the mass $G + g$ (which includes the piston itself) and an atmospheric pressure = $A$, constitute a constant back-pressure. Through the portion $OB$ of stroke $s_o$, of the steam-piston, the steam works while beyond $B$, boiler communication being "cut off." $S$ diminishes by Boyle's law, i.e., in this case, for any point above $B$, we have, neglecting the "clearance," $F$ being the cross-section of the cylinder, $$S : S_1 : F_1 : F_2 ; \text{ or } S = S_{A_r} + s_o.$$ Full length of stroke = $ON = s_o$. Given, then, the forces $S$ and $A$, the distances $s$ and $s_o$ and the velocities at $O$ and at $N$ both = 0 (i.e., the mass $M = G + g$ is to start from rest at $O$, and to come to rest at $N$), required the proper weight $G$ to 62 MECHANICS OF ENGINEERING. fulfil these conditions, $S$ varying as already stated. The acceleration at any point will be $$p = (S - A - G) + M \quad . . . . . . . . . . . . . . . . . (1)$$ Hence (eq. III.) $Mds = (S - A - G)ds$, and $\cdot\cdot\cdot$ for the whole stroke $$M \int_0^{\infty} vds = \int_0^{\infty} (S - A - G)ds; \text{ i.e.,}$$ $$0 = Ss \int_0^{\infty} ds + Ss \int_0^{\infty} \frac{ds}{s} - A \int_0^{\infty} ds - G \int_0^{\infty} ds,$$ or $$Ss \left[ 1 + \log \frac{s}{s_0} \right] = As + Gs \quad . . . . . . . . . . (2)$$ Since $S = S_0$ is constant, from $O$ to $B$, and variable, $= Ss$, $\cdot s$, from $B$ to $N$, we have had to write the summation $\int_0^{\infty} Sds$ in two parts. From (2), $G$ becomes known, and $\cdot\cdot\cdot M$ also ($= G + g$). Required, further, the time occupied in this upward stroke. From $O$ to $B$ (the point of entry) the motion is uniformly accelerated, since $p$ is constant ($S$ being $S_0$, in eq. (1)), with the initial velocity $v$. Hence, by eq. (3), the velocity at $B$ is $v = \sqrt{2(S_0 - A - G)}$, and $M$ is known; $\cdot\cdot\cdot$ the time $t_B = \frac{v}{g}$, becomes known (eq. (4), p. 56) of describing OB. At any point beyond $B$ the velocity may be obtained thus: From (III.) $vds = pds$, and eq. (1) we have, summing between $B$ and any point above, $$M \int_{B_n}^{\infty} vds = Ss \int_{B_n}^{\infty} \frac{ds}{s} - (A + G) \int_{B_n}^{\infty} ds; \text{ i.e.,}$$ $$\frac{G(v^2 - v_0^2)}{2} = Ss, \log_s \frac{s}{s_n} - (A + G)(s - s_n).$$ This gives the relation between the two variables $v$ and $s$ anywhere between $B$ and $N$, if we solve for $v$ and insert its value in $dt = ds + dv$, we shall have that $dt$ is a function of $s$ and $ds$, which is not integrable. Hence we may resort to approxi- RECTILINEAR MOTION OF A MATERIAL POINT. 68 mate methods for the time from $B$ to $N$. Divide the space $BN$ into an uneven number of equal parts, say five; the dis- tances of the points of division from $O$ will be $s_0$, $s_0 + s_1$, $s_0 + s_2$, and $s_0$. For each value of $s$ compute (from above) the velocity $v_s$ (already known), $v_{s-1}$, $v_{s-2}$, $v_{s-3}$ and $v_s$ (known to be zero). To the first four spaces apply Simpson's Rule, and we have the time from $B$ to the end of $s_4$ $$\int_{s_0}^{s_4} \frac{ds}{v_s}; \text{ approx. } = \frac{s_4 - s_0}{12}\left[\frac{4}{v_4} + \frac{2}{v_3} + \frac{4}{v_2} + \frac{1}{v_1}\right];$$ while regarding the motion from $5$ to $N$ as uniformly retarded (approximately) with initial velocity = $v_s$ and the final = zero, we have (eq. (4), § 56), $$\int_{s_5}^{s_N} t = 2(s_n - s_4)/v_s.$$ By adding the three times now found we have the whole time of ascent. In Fig. 67 the dotted curve on the left shows by horizontal ordinates the variation in the velocity as the distance $s$ increases; similarly on the right are ordinates showing the variation in the pressure as the distance $s$ increases. The maximum force, being $\sigma v_s$, may be found by putting $p = 0$, i.e., for $S = M + G$, the accelerating force being = 0, see eq. (1). Below $E$ the ac- celerating force, and consequently the acceleration, is positive; above, negative (i.e., the back-pressure exceeds the steam- pressure). The horizontal ordinates between the line HEKL and the right line RT are proportional to the accelerating force. If by condensation of the steam a vacuum is produced be- low the piston on its arrival at $N$, the accelerating force is downwards. It is left to the student determine how the detail of this problem would be changed, if the cylinder were horizontal instead of vertical.) 60. Direct Central Impact.—Suppose two masses $M_1$ and $M_2$, to be moving in the same right line so that their distance apart, continually diminishes, and that when the collision or impact takes place the line of action of the mutual pressure coincides with the line joining their centres of gravity, or centres of 64 MECHANICS OF ENGINEERING. mass, as they may be called in this connection. This is called a direct central impact, and the motion of each mass is variably accelerated and rectilinear during their contact, the only force being the pressure of the other body. The whole mass of each body will be considered concentrated in the centre of mass, on the supposition that all its particles undergo simultaneously the same change of motion in parallel directions. (This is not strictly true; the effect of the pressure being gradually distributed over the surface of the bodies diminishes and at some extent after the impact.) When the centres of mass cease to approach each other the pressure between the bodies is a maximum and the bodies have a common velocity; after this, if any capacity for restitution of form (elasticity) exists in either body, the pressure still continues, but diminishes in value gradually to zero, when contact ceases and the bodies separate with different velocities. Reckoning the time from the first instant of contact, let $t$ be duration of the first period; just before this instant let $v_1$ be the final velocity (restitution). Fig. 88. Let $M_1$ and $M_2$ be the masses, and at any instant during the contact let $v_1$ and $v_2$ be simultaneous values of the velocities of the mass-centres respectively (reckoning velocities positive toward the right), and $P$ the pressure (variable). At any instant the acceleration of $M_1$ is $p_1 = P - M_1$, while at the same instant that of $M_2$ is $p_2 = P + M_2$. $M_1$ being retarded, $M_2$ accelerated, in velocity. Hence (eq. II., p. 40 $\rightarrow$ p. 43) we have $$M_1 v_1 dt = - Pdt; \quad M_2 v_2 dt = + Pdt \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ . \ .\ (1)$$ Summing all similar terms for the first period of the impact, we have (calling the velocities before impact $c$, and $c_p$ and the common velocity at instant of maximum pressure $C$) $$M_1\int_{t_0}^{t} dv_1 = -\int_{t_0}^{t} Pdt, i.e., M_1(C-c_1) = -\int_{t_0}^{t} Pdt; \\ (9)$$ $$M_2\int_{t_0}^{t} dv_2 = +\int_{t_0}^{t} Pdt, i.e., M_2(C-c_2) = +\int_{t_0}^{t} Pdt. \\ (3)$$ RECTILINEAR MOTION OF A MATERIAL POINT. 65 The two integrals are identical, numerically, term by term, since the pressure which at any instant accelerates $M_t$ is numerically equal to that which retards $M_t$; hence, though we do not know how the pressure varies with time, the two integrals are equal. We have then a definite interval between (9) and (8) and solve for $C$. If the impact is *inelastic* (i.e., no power of restitution in either body, either on account of their total inelasticity or damaging effect of the pressure at the surfaces of contact), they continue to move with this common velocity, which is therefore their final velocity. Solving, we have $$C = \frac{M_{c_1} + M_{c_2}}{M_t + M_s} \quad \ldots \quad (4)$$ Next, supposing that the impact is *partially elastic*, that the bodies are of the same material, and that the summation $\int_{t_i}^{t_f} Pdt$ for the second period of the impact bears a ratio, $e$, to that $\int_{t_i}^{t_f} Pdt$, already used, a ratio peculiar to the material, if the impact is not too severe, we have, summing equations (1) for the second period (letting $V_c$ and $V_s$ be the velocities after impact), $$M_t \int_{t_i}^{t_f} dv_t = -\int_{t_i}^{t_f} Pdt; \text{i.e., } M_t(V_s - C) = -e \int_{t_i}^{t_f} Pdt; \quad (5)$$ $$M_s \int_{t_i}^{t_f} dv_s = +\int_{t_i}^{t_f} Pdt; \text{i.e., } M_s(V_s - C) = +e \int_{t_i}^{t_f} Pdt. \quad (6)$$ $e$ is called the coefficient of restitution. Having determined the value of $\int_{t_i}^{t_f} Pdt$ from (9) and (8) in terms of the masses and initial velocities, substitute it and (7) from (4), in (5), and we have (for the final velocities) $$V_c = [M_{c_1} + M_{c_2} - e(M_t(c_t - c_0)] + [M_t + M_s]; \quad (7)$$ and similarly $$V_s = [M_{c_1} + M_{c_2} + e(M_t(c_t - c_0)] + [M_t + M_s]. \quad (8)$$ For $e = 0$, i.e., for inelastic impact, $V_c = V_s = C$ in eq. (4); for 68 MECHANICS OF ENGINEERING. $e = 1$, or *elastic impact*, (7) and (8) become somewhat simplified. To determine experimentally, let a ball ($M_1$) of the substance fall upon a very large slab ($M_2$) of the same substance, noting both the height of fall $h_1$, and the height of rebound $H_2$. Considering $M_1$ as $\alpha M_2$, with $$c_1 = \sqrt{2gh_1}, V_1 = -\sqrt{2gh_2}$$ and $c_2 = o$, eq. (7) gives $$-\sqrt{2gh_2} = -e\sqrt{2gh_1}, e : e = H_2 : h_1$$ Let the student prove the following from equations (9), (3), (5), and (6). (a) For any direct central impact whatever, $$M_{c_1} + M_{c_2} = M_1 + M_2, V_{c_1} = V_{c_2}$$ [The product of a mass by its velocity being sometimes called its momentum, this result may be stated thus: In any direct central impact the sum of the momenta before impact is equal to that after impact (or at any instant during impact). This principle is called the Conservation of Momentum. The present is only a particular case of a more general proposition.] It may be proved $C$, eq. (4), is the velocity of the centre of gravity of the two masses before impact; the conservation of momentum, then, asserts that this velocity is unchanged by the impact, i.e., by the mutual actions of the two bodies.] (b) The loss of velocity of $M_1$, and the gain of velocity of $M_2$, are twice as great when the impact is elastic when inelastic. (c) If $e = 1$, and $M_1 = M_2$, then $V_1 = c_1$, and $V_2 = c_2$. Example—Let $M_1$ and $M_2$ be perfectly elastic, having weights $= 4$ and 5 lbs. respectively, and let $e = 10$ ft. per sec, and $c_1 = 6$ ft. per sec. (i.e., before impact $M_1$ is moving in a direction contrary to that of $M_2$). By substituting in eqs. (7) and (8), with $e = 1$, $M_1 = 4 + g$, and $M_2 = 5 + p$, we have $$V_1 = \frac{4}{g}(4 \times 10 + 5 \times (-6) - 5(10 - (-6))) = -7.7 \text{ ft. per sec.}$$ $$V_2 = \frac{5}{g}(4 \times 10 + 5 \times (-6) + e(10 - (-6))) = +8.2 \text{ ft. per sec.}$$ as the velocities after impact. Notice their directions, as indicated by their signs "VIRTUAL VELOCITIES." 67 CHAPTER II. "VIRTUAL VELOCITIES." 61. Definition. —If a material point is moving in a direction not coincident with that of the resultant force acting (as in curvilinear motion in the next chapter), and any element of its path, $ds$, projected upon this force, *the length of this projection, $ds$, Fig. 69, is called the "Virtual Velocity" of the force, since $ds + dt$ may be considered the velocity of the force at this instant, just as $ds + dt$ is $ds + dt$ that of the point. The product of the force by $\frac{ds}{dt}$ will be called its virtual moment, reckoned + or - according as the direction from $O$ to $D$ is the same as that of the force or opposite. Fig. 68 68 62. Prop. I.—The virtual moment of a force equals the algebraic sum of those of its components. Fig. 70. Take the direction of $ds$ as an axis $x$; let $P$ and $P_1$ be components of $F$; $\sigma_x$, $\sigma_y$, and $\alpha$ their angles with $x$. Then (§ 16) $P \cos \alpha = P_x \cos \alpha + P_y \cos y$. Hence $P(d\sigma \cos \alpha)$ = $(P_x ds \cos \alpha + P_y ds \cos y) = P ds (\cos \alpha + \cos y)$ = the projection of $ds$ upon $P$, i.e., = $du$; $\therefore ds \cos \alpha = du$, etc.; $\therefore Pdu = P_x du + P_y du$. If in Fig. 70, $\sigma_x$ were > 90°, evidently we would have $Pdu = -P_x du + P_y du$, i.e., $Pdu$ would then be negative, and $OD$ would fall behind $O$; hence the definition of + and - in § 61. For any number of components the proof would be similar, and is equally applicable whether they are in one plane or not. 63. Prop. II.—The sum of the virtual moments equals zero for concurrent forces in equilibrium. * We should rather say " projected on the line of action of the force; but the phrase used may be allowed, for brevity. Fig. 70 68 MECHANICS OF ENGINEERING. (If the forces are balanced, the material point is moving in a straight line if moving at all.) The resultant force is zero. Hence, from § 62, $P_du + P_du + \ldots = 0$, having proper regard to sign, i.e., $Z(Pdu) = 0$. **Prop. III.** —The sum of the virtual moments equals zero for any small displacement or motion of a rigid body in equilibrium under non-concurrent forces in a plane; all points of the body moving parallel to this plane. (Although the kinds of motion of a given rigid body which are consistent with balanced non-concurrent forces have not yet been investigated, we may imagine any slight motion for the sake of the algebraic relations between the different $du$'s and forces.) First, let the motion be a translation, all points of the body being parallel to equal parallel lengths $= du$. Take $X$ parallel to $du$; let $\alpha$, etc., be the angles of the forces with $X$. Then ($\S 35$) $\Sigma Z(P\cos\alpha) = 0$; $\Sigma Z(P\cos\alpha)$ $= 0$ but $de\cos\alpha = du$; $de\cos\alpha = du$; etc.; $\Sigma Z(Pdu) = 0$. Q. E. D. Secondly, let the motion be a rotation through a small angle $d\theta$ in the plane of the forces about any point $O$ in that plane, Fig. 72. With $O$ as a pole let $\rho_1$ be the radius-vector of the application of $F_1$, and $a$, its lever-arm from $O$; similarly for the other forces. In the rotation each point of application describes a small arc, $de_1$, $de_2$, etc., proportional to $\rho_1$, $\rho_2$, etc., since $de_1 = \rho_1 d\theta$, $de_2 = \rho_2 d\theta$, etc. From § 86, $F_1a_1 + \ldots + F_n a_n = 0$ but from similar triangles $de_1 : de_2 : \ldots : a_1 = gdu : gdu : \ldots = gdu : gdu : \ldots = du : du : \ldots$ similarly $a_1 : du : \ldots = J : d\theta : \ldots$. Hence we must have $P_1du + P_2du + \ldots + J + d\theta = 0$, i.e., $Z(Pdu) = 0$. Q. E. D. Since any slight displacement or motion of a body may be conceived to be accomplished by a small translation followed by a rotation through a small angle, and since the fore- "VIRTUAL VELOCITIES." going deals only with projections of paths, the proposition is established and is called the Principle of Virtual Velocities. [A similar proof may be used for any slight motion whatever in space when a system of non-concurrent forces is balanced.] Evidently if the path ($ds$) of a point of application is perpendicular to the direction of the velocity ($dv$), consequently the virtual moment ($Pdv$) of the force is zero. Hence we may frequently make the displacement of such a character in a problem that one or more of the forces may be excluded from the summation of virtual moments. 65. Connecting Rod by Virtual Velocities.—Let the effective steam-pressure $P$ be the means, through the connecting-rod and crank (i.e., two links), of raising the weight $G$ very slowly; neglect friction and the weight of the links themselves. Consider $AB$ as free (see $\theta$ in Fig. 78); $BC$ also, at $\theta$; let the A diagram showing a connecting rod mechanism with points A, B, C, D labeled. (a). (b). (c). "small displacements" of both be simultaneous small portions of their ordinary motion in the apparatus. $A$ has moved to $A_1$, through $da$; $B$ to $B_1$, through $da$; a small arc; $O$ has not moved. The forces acting on $AB$ are $P$ (steam-pressure), $N$ (vertical reaction of guide), and $N'$ and $T$ (the tangential and normal components of the crank-pin pressure). Those on $BC$ are $N'$ and $T$ (reversed), the weight $G$, and the oblique pressure of bearing $P'$. The motion being slow (or rather the acceleration being small), each of these two systems will be considered as balanced. Now put $(Fda) = 0$ for $AB$, and we have $$Pda + N' \times 0 + N' \times 0 - Tda = 0.$$ . . . (1) For the simultaneous and corresponding motion of $BC$, $(E(Pda)) = 0$ gives TO MECHANICS OF ENGINEERING. $N' \times 0 + Tds - Gdh + P' \times 0 = 0, \quad (3)$ $dh$ being the vertical projection of $G$ of motion. From (1) and (3) we have, easily, $Pda - Gdh = 0, \quad (3)$ which is the same as we might have obtained by putting $2(Pda) = 0$ for the two links together, regarded collectively as a free body, and describing a small portion of the motion they really have in the mechanism, viz., (Fig. 74.) $Pda + N' \times 0 - Gdh + P' \times 0 = 0. \quad (4)$ We may therefore announce the— 66. Generality of the Principle of Virtual Velocities.—If any mechanism of flexible inextensible cords, or of rigid bodies jointed together, or both, at rest, or in motion with very small accelerations, be considered free collectively (or any portion of it), and all the external forces put in; then (disregarding mutual frictions), for a small portion of its prescribed motion, $2(Pda)$ must $= 0$, in which the $du$, or virtual velocity, of each force, $P$, is the projection of the path of the point of application upon the force (the product, $Pda$, being $+or-$ according to § 61). 67. Example.—In the problem of § 65, having given the weight $G$, required the proper steam-pressure (effective) $P$ to hold $G$ in equilibrium, or to raise it uniformly, if already in motion, for a given position of the links. That is, Fig. 75, given $a$, $r$, $c$, $\alpha$, and $\beta$, required the ratio $dh : de$; for, from equation (3), § 65, $= Gdh : Pde = AB : AC$ projections of $de$ and $ds$ upon $AB$ will be equal, since $AB = A'B'$, and makes an (infinitely) small angle with $A'B'$, i.e., $dc\cos\alpha = de\cos(\beta-\alpha)$. Also, $dh = (e : r)de\sin\beta$ Fig. 74. Fig. 75. "VIRTUAL VELOCITIES." Eliminating $ds$, we have, $$\frac{dh}{dx} = \frac{c \sin \beta \cos \alpha}{\cos(\beta - \alpha)}; \quad P = G \frac{\sin \beta \cos \alpha}{\cos(\beta - \alpha)}$$ 68. When the acceleration of the parts of the mechanism is not practically zero, $\mathcal{Z}(Pdu)$ will not be 0, but a function of the masses and velocities to be explained in the chapter on Work, Energy, and Power. If friction occurs at moving joints, enough "free bodies" should be considered that no free body extends beyond each joint; it will be found that this friction cannot be eliminated in the way in which $T$ and $N$ were, in § 65. 69. Additional Problems: to be solved by "virtual velocities." Problem 1.--Find relations between the forces acting on a straight lever in equilibrium; also, on a bent lever. Example 2. When an ordinary screw-eyewhisk is in equilibrium, find the relation between the force applied horizontally, and tangentially at the circumference of the wheel, and the vertical resistance under the screw-eyewhisk. Solution.--Considering free the rigid body consisting of the wheel and screw-eywhisk, let $R$ be the resistance at the point of shaft (pointing along the line of action of the force) of the wheel, and $r$ the radius of edge of wheel. Let radius of wheel be $r$. Reaction $A$ and $P$ these also acting on this body certain pressures, or "supporting forces," consisting of the reactions of the collar, and reactions of the threads of nut against the thread of screw-eywhisk. The reaction $A$ being equal to the resistance $R$, the shaft would advance for a full turn of the wheel. Then if we imagine the wheel to turn through a small angle $\theta$, the corresponding advance, $ds$, of the shaft would be from the proportion $ds : ds = 3r : R$. The path of the point of application of $P$ would be a small portion of a helix, the projection of which on the line of $P$ is $r\theta$, while its distance is its full length on the line of the force. In the case of each of the other forces, their projections on this line are proportional to their distances from this line (which is normal to the rubbing surface), friction being disregarded). Hence, substituting in $\mathcal{Z}(Pdu) = 0$, we have $$+ F.r\theta - H.dh + 0 + 0 = 0$$ whence $$F = r\theta.R = \frac{e}{2\pi}.R$$ 73 MECHANICS OF ENGINEERING. CHAPTER III. CURVILINEAR MOTION OF A MATERIAL POINT. [Motion in a plane, only, will be considered in this chapter.] 70. Parallelogram of Motiona.—It is convenient to regard the curvilinear motion of a point in a plane as compounded, or made up, of two independent rectilinear motions parallel respectively to two co-ordinate axes $X$ and $Y$, which may be expressed thus: Fig. 76. Consider the drawing-board $CD$ as fixed, and let the head of a $T$-square move from $A$ toward $B$ along the edge according to any law whatever, while a pencil moves from $M$ toward $Q$ along the blade. The result is a curved line on the board, whose form depends on the character of the two $X$ and $Y$ component motions, and is therefore called. If at time $t_1$, the $T$-square head moves an $X$ distance $MN$, and the pencil simultaneously a $Y$ distance $MP$, by completing the parallelogram on these lines, we obtain $R_t$, the position of the point on the board at the end of the time $t_1$. Similarly, at the end of the time $t_2$, we find the point at $E'$. 71. Parallelogram of Velocities.—Let the $X$ and $Y$ motions be uniform, required the resulting motion. Fig. 77. Let $\vec{c}_x$ and $\vec{c}_y$ be the constant uniform $X$ and $Y$ velocities. Then in any time interval $\Delta t$, i.e., when we have, eliminating $t$, $$x + y = \vec{c}_x + \vec{c}_y = \text{constant},$$ i.e., $x$ is proportional to $y$, i.e., the path is a straight line. Laying off $OA = \vec{c}_x$, and $AB = \vec{c}_y$, $B$ is a point of the path, and $OD$ is the distance described by the point in the first Fig. 76. Fig. 77. CURVILINEAR MOTION OF A MATERIAL POINT. 73 second. Since by similar triangles $OR : x = OH : o_p$, we have also $OR = OB \cdot t$; hence the resultant motion is uniform, and its velocity, $OB = c$, is the diagonal of the parallelogram formed on the two component velocities. Corollary.--If the resultant motion is curved, the direction and velocity of the motion at any point will be given by the diagonal formed on the component velocities at that instant. The direction of motion is, of course, a tangent to the curve. **72. Uniformly Accelerated X and Y Motion.—The initial velocities of both being zero.** Required the resultant motion. Fig. 78. From §56, eq. (8) (both $o_x$ and $o_y$ being $=0$), we have $a = \frac{1}{2}p_x p_y$ and $y = \frac{1}{2}p_y^2$, whence $x + y = p_x + p_y$ constant, and the resultant motion is in a straight line. Consecutive lines laid off from $O$ on $x$ and $Y$ to represent $p_x$ and $p_y$ to scale will form a straight line. Then similar triangles ($OR$ being the distance described in the resultant motion in any time), $OB : x = OR : p_x \therefore OR = 4tOB$. Hence, from the form of this last equation, the resultant motion is uniformly accelerated, and its acceleration is $OB = p_x$ (on the same scale as $p_x$ and $p_y$). This might be called the parallelogram of accelerations, but is really a parallelogram of forces, if we consider that a free material point, initially at rest at $O$, and acted on simultaneously by constant forces $P_x$ and $P_y$ (so that $p_x = P_x + M$ and $p_y = P_y + M$), must begin a uniformly accelerated rectilinear motion in the direction of the resultant force, having no initial velocity in any direction. **73. In general, considering the point hitherto spoken of as a free material point, under the action of one or more forces, in view of the foregoing, and of Newton's second law, given the initial velocity in amount and direction, the starting-point, the initial amounts and directions of the acting forces and the** A diagram showing a free body O with forces P_x and P_y applied at points P_x' and P_y'. The diagram shows how these forces can be resolved into components P_x' and P_y', which act along lines OP_x' and OP_y', respectively. 74 MECHANICS OF ENGINEERING. laws of their variation if they are not constant, we can resolve them into a single $X$ and a single $Y$ force at any instant, determine the $X$ and $Y$ motions independently, and afterwards the resultant motion. The resultant force is never in the direction of the tangent to the path at any point (see section 6). The relations which it amounts to and direct towards at any instant bear to the velocity, the rate of change of that velocity, and the radius of curvature of the path will appear in the next paragraph. 74. General Equations for the curvilinear motion of a material point in a plane.—The motion will be considered resulting from the composition of independent $X$ and $Y$ motions, $X$ and $Y$ being perpendicular to each other. Fig. 79. In two consecutive equal times (each = $dt$), let $dx$ and $dy$ = small spaces due to the $X$ motion; and $dy$ and $dK = dy$, due to the $Y$ motion. Then $dx$ and $dy$ are two consecutive elements of long $de$, making $BE = de$; the $EF = dx$, $CF = dy$, and $CO = d'x (EO being perpendicular to BE). Also draw CL perpendicular to BG and call CL $d'n$. Call the velocity of the $X$ motion $v_x$, its acceleration $p_x$; those of the $Y$ motion, $v_y$ and $p_y$. Then, $$v_x = \frac{dx}{dt}; v_y = \frac{dy}{dt}; p_x = \frac{dv_x}{dt} = \frac{d^2x}{dt^2}; \text{and } p_y = \frac{dv_y}{dt} = \frac{d^2y}{dt^2}. $$ For the velocity along the curve (i.e., tangent) $$v = dx + dy, \text{we shall have, since } dv = dx + dy.$$ Hence $$v' = \left(\frac{dx}{dt}\right)' + \left(\frac{dy}{dt}\right)' = v_x' + v_y'. \quad (1)$$ Let $p_t$ be the acceleration of $v$, i.e., the tangential acceleration. CURVILINEAR MOTION OF A MATERIAL POINT. 75 then $p_1 = d^2s + dt^2$, and, since $d^2s$ is the sum of the projections of $EF$ and $CF$ on $BC$, i.e., $d^2s = d^2x \cos \alpha + d^2y \sin \alpha$, we have $$\frac{d^2s}{dt^2} = \frac{d^2x}{dt^2} \cos \alpha + \frac{d^2y}{dt^2} \sin \alpha; \text{ i.e., } p_1 = p_x \cos \alpha + p_y \sin \alpha. (9)$$ By Normal Acceleration we mean the rate of change of the velocity in the direction of the normal. In describing the element $AB = ds$, no progress has been made in the direction of the normal $BH$ i.e., there is no velocity in the direction of the normal. But, as long as the point describes a portion of path, the point has progressed a distance $CL$ (call it $dn$) in the direction of the old normal $BH$ (though none in that of the new normal $CI$). Hence, just as the tang. acc. $$= \frac{ds - ds}{dt} = \frac{d^2x}{dt^2},$$ so the normal acc. = $\frac{CL - zero}{dt} = \frac{d^2n}{dt^2}$. It now remains to express this normal acceleration ($= p_n$) in terms of the $X$ and $Y$ accelerations. From the figure, $CL$ $$= CM - ML,$$ i.e., $$d^2n = d^2y \cos \alpha - d^2x \sin \alpha;$$ since $EF = d^2x$; $$\therefore \frac{d^2n}{dt^2} = \frac{d^2y}{dt^2} \cos \alpha - \frac{d^2x}{dt^2} \sin \alpha.$$ Hence $$p_n = p_x \cos \alpha - p_y \sin \alpha.$$ (3) The norm. acc. may also be expressed in terms of the tang. velocity $v$, and the radius of curvature $r$, as follows: $$ds = rda; \text{ or } da = dv/r;$$ also $$d^2n = d^2da/dt^2 = dv/dt + r,$$ i.e., $$\frac{d^2n}{dt^2} = (\frac{dv}{dt})\frac{1}{r}$$ or $$p_n = v\frac{v}{r}. (4)$$ If now, Fig. 80, we resolve the forces $X = M_p x$ and $Y = M_p y$ which at this instant account for the X and Y accelerations ($M =$ mass of the material point), into components along the tangent and normal to the curved path, shall have, as their equivalent, a tangential force A diagram showing a curved path with tangents drawn at various points. $$T = M_p x + M_p y \sin \alpha.$$ 76 MECHANICS OF ENGINEERING. and a normal force $$N = M_p \cos \alpha - M_p \sin \alpha.$$ But, see equation (5), similarly we may also write $$T = M_p + \frac{d^2y}{dt^2}$$ and $$S = M_p = \frac{dy}{dt}.$$ (5) Hence, if a free material point is moving in a curved path, the sum of the tangential components of the acting forces must equal the mass $\times$ tangent, since this is a real component, = the mass $\times$ normal, since = the mass $\times$ square of velocity in path, $+$ real curve. It is evident, therefore, that the resultant force (= diagonal on $T$ and $N$) nor $T$ and $N$, does not act along the tangent at any point but toward the concave side of the path; namely, $$r = x.$$ **Radius of curvature.** From the line above eq. (4) we have $ds = dx^2 + r$; hence line above eq. (3), $ds + r = dy\cos \alpha - dy\sin \alpha$; because $ds = dx + ds$, and sin $\alpha = dy/dx$, i.e., $$ds/r = d^2y/dt^2 - d^2y/dx^2$$ or $$r = d^2y/dt^2 - d^2y/dx^2$$ i.e., $$\frac{ds}{r} = d^2y/dt^2 - d^2y/dx^2$$ or $$r = \sqrt{\frac{ds}{dt}} - \frac{d^2y}{dx^2}\cdot\frac{d\tan \alpha}{dt}$$ or, $$r = e^{\prime} + \left(e_{\alpha}^{'}\cdot\frac{d\tan \alpha}{dt}\right)$$ (6) which is equally true if, for $e_{\alpha}$ and tan $\alpha$, we put $e_{\alpha}$ and tan $(0^\circ - \alpha)$ respectively. **Change in the velocity square.** Since the tangential acceleration $\frac{ds}{dt} = p_0$, we have $\frac{ds}{dt} = p_0 dt$; i.e., $$ds/p_0 = p_0 dt,$$ or $vdo = p_0 dt$ and $\therefore \frac{v^2 - v_0^2}{2} = f p_0 da.$ (7) having integrated between any initial point of the curve where $v = c$, and any other point where $v = v$. This is nothing more than equation (III.), of § 50. CURVILINEAR MOTION OF A MATERIAL POINT. 77 75. Normal Acceleration. Second Method.—Fig. 81. Let $C$ be the centre of curvature and $OD = 2r$. Let $OB'$ be a portion of the osculating parabola (vertex at $O$; any osculating curve will serve). When $ds$ is described, the distance passed over in the direction of the normal $AB$; for $2ds$, it would be $2AB' = AB + AB''$ (this square of $OB'$; property of a parabola), and so on. Hence the motion along the normal is uniformly accelerated with initial velocity = 0, since the distance $AB$, varies as the square of the time (considering the motion along the curve of uniform velocity, so that the distance $OB$ is directly as the time). If $p_s$ denote the accel of this uniformly accelerated motion, its initial velocity being 0, we have (eq. 2, § 6) $AB' = p_s ds$, i.e., $p_s = \frac{d^2s}{dt^2} ds$. But from the similar triangles $\triangle OAB$ and $\triangle OAB'$, we have $\frac{AD}{AB'} = \frac{ds}{dt}$, hence $2AB' = ds + r$, i.e., $p_s = \frac{ds}{dt} + r$, hence $p_s = \frac{ds}{dt} - r$. 76. Uniform Circular Motion. Centripetal Force.—The velocity being constant, $p_s$ must be 0, and $T$ (or $2T$ if there are several forces) must be 0. The resultant of all the forces, therefore, must be a normal force $(Mv^2 + r)$ = a constant (eq. 5, § 74). This is called the "deviating force," or "centripetal force;" without it the body would continue in a straight line. Since forces always occur in pairs ($\S 3$), a "centrifugal force," equal and opposite to the "centri- petal" (one being the reaction of the other), will be found among the forces acting on the body which constrain the deviation of the first body from its natural straight course is due. For example, the attraction of the earth on the moon acts as a centripetal or deviating force on the latter, while the equal and opposite force acting on the earth may be called the centrifugal. If a small block moving on a smooth horizontal table is gradually turned from its straight course $AB$ by a fixed circular guide, tangent to $AB$ at $B$, the pressure of the guide will act with a centripetal or deviating force $Mv^2/r$ directed toward the centre of curvature, while A diagram showing a block moving on a smooth horizontal table and being gradually turned from its straight course AB by a fixed circular guide tangent to AB at B. 78 MECHANICS OF ENGINEERING. the centrifugal force $Mv + r$ is the pressure of the block against the guide, directed away from that centre. The cen- trifugal forces, then, is never found among the forces acting on the body, but only as a reaction to its motion with The Conical Pendulum, or governor-ball.—Fig. 82. If a material point of mass $= M = G + g$, suspended on a cord of length $l$, is to maintain a uniform cir- cular motion in a horizontal plane, with a given radius $r$, under the action of gravity and the cord, required the velocity $v$ to be given by $v^2 = G + g$. At $B$ we have an ideal body free. For the only force acting on $G$ is the cord- tension $T$. Thus, of their normal com- ponents, i.e., $N$, must $= Mg + r$, i.e., $P \sin \alpha = Mg + r$; but, since $S$ (vert. comp.) = 0, $P \cos \alpha = G$. Hence $G \tan \alpha = G + gr$; i.e., $\alpha = \sqrt{g} \tan \alpha$. Let $u =$ number of revolutions per unit of time, then $u = c + 2\pi r = v/g + 2\pi \sqrt{A}$; i.e., is inversely proportional to the (vertical projection) of the cord-length. The time of one revolution is $t = 1/u$. Elevation of the outer rail on railroad curves (considera- tions of traction disregarded).—Consider a single car as a material point, and free, having a given velocity $v$ and being subject to a vertical pressure against the wheels. So long as the car follows the track the resultant $R$ of $P$ and $Q$ must point toward the centre of curvature and have a value of $Mc + r$. But $R = G \tan \alpha$, whence $\tan \alpha = c^2 + gr$. If therefore the ties are placed at this angle $\alpha$ with the horizontal, the pressure will come upon the tread and not on the flanges of the wheels ; in other words, the car will follow the track. (This is really the same problem as the preceding.) Apparent weight of a body at the equator.—This is less than the true weight or attraction of the earth, on account of the uniform circular motion of the body with the earth in its diurnal rotation. If the body hangs from a spring-balance, A diagram showing a pendulum-like structure with a central point labeled "G" connected to two points labeled "P" and "Q". The angle between "P" and "Q" is labeled as "α". A line from "G" to "Q" forms an angle with "Q" equal to "α". Another line from "G" to "P" forms an angle with "P" equal to "α". The distance between "P" and "Q" is labeled as "c". The distance between "G" and "Q" is labeled as "r". The distance between "G" and "P" is labeled as "g". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as "α". The angle between "G", "Q", and "P" is labeled as CURVILINEAR MOTION OF A MATERIAL POINT. 76 whose indication is $G$ lbs. (apparent weight), while the true attraction is $G'$ lbs, we have $G' - G = M\alpha - \tau$. For $M$ we may use $G + \tau$ (apparent value), for $r$ about 90,000,000 ft.; for $c$, 25,000 miles in 24 hrs., reduced to feet per second. It results from this that $G$ is $< G'$ by $\frac{1}{24} G r$ nearly, and (since $17^{\circ} = 289$) hence if the earth revolved on its axis seven-teen times as fast as at present, $G$ would = 0, i.e., bodies would apparently have no weight, the earth's attraction on them being just equal to the necessary centripetal or deviating force necessary to keep the body in its orbit. Certainly the earth was originally a liquid, and at rest, its form would be spheroidal; but when revolving uniformly around the polar diameter, its form of relative equilibrium (i.e., no motion of the particles relatively to each other) is nearly ellipsoidal, the polar diameter being an axis of symmetry. Lines of attraction on bodies at its surface do not intersect in a common point, and the centripetal force requisite to keep a suspended body in its orbit (a small circle of the ellipsoid), at any point on its surface, is directed toward the centre of apparent weight $G$ directed (nearly) toward the centre, and of $G$ the tension of the string. Fig. 85. $G = \text{the apparent weight},$ indicated by a spring-balance and $MA$ is its line of action (plumb-line) normal to the ocean surface. Evidently the apparent weight, and consequently $g$, are less than the true values, since $N$ must be perpendicular to the polar axis, while the true value is parallel to it. Hence, as $r$ increases, the square of $MC$, decreases toward the equator, hence the apparent values decrease still more rapidly as the latitude diminishes. The following equation gives the apparent $g$ for any latitude $\beta$, very nearly (units, foot and second): $$g = 32.1508 - 0.0821 \cos 2\beta.$$ (The value 32.15 is accurate enough for practical purposes. Since the earth's axis is really not at rest, but moving about) A diagram showing lines of attraction on a body at the surface of a rotating sphere. 80 MECHANICS OF ENGINEERING. the sun, and also about the centre of gravity of the moon and earth, the form of the ocean surface is periodically varied, i.e., the phenomena of the tides are produced. 77. Cycloidal Pendulum.—This consists of a material point at the extremity of an impenetrable, flexible, and inextensible cord of length = $l$, confined to the arc of a cycloid in a vertical plane by the cycloidal evolute shown in Fig. 86. Let the oscillation begin (from rest) at $A$, a height = $h$ above $O$ A diagram showing a cycloidal pendulum with a point $P$ at the end of a flexible, inextensible cord following a cycloid path. Fig. 86 the vertex. On reaching any lower point, as $B$ (height = $s$ above $O$), the point has acquired some velocity which is at this instant increasing at some rate = $\frac{ds}{dt}$. Now consider the point free, Fig. 87; the forces acting are $P$ the cord-tension, normal to path, and $G$ the weight, at an angle $\varphi$ with the path. From § 74, eq. (5), $2T = Mg$, gives $$G \cos \varphi + P \cos 90^\circ = (G + p)g; \therefore p = g \cos \varphi$$ Hence (eq. (7), § 74), $evd\varphi = pds$ gives $$evd\varphi = g \cos \varphi ds; but ds \cos \varphi = - dx; \therefore evd\varphi = - gdx.$$ Summing between $A$ and $B$, we have $$\int_{A}^{B} evd\varphi = - g \int_{A}^{B} dx; \text{or } v^2 = 2g(h-s);$$ the same as if it had fallen freely from rest through the height $h-s$. (This result evidently applies to any form of path when, besides the weight $G$, there is but one other force, and that always normal to the path.) From $2N = Me^2 + r_1$, we have $P - G \sin \varphi = Mo^2 + r_1,$ CURVILINEAR MOTION OF A MATERIAL POINT. 81 whence $P$, the cord-tension at any point, may be found (here $\pi$=the radius of curvature at any point = length of straight portion of the cord). To find the time of passing from $A$ to $O$, i.e., half-oscillation, substitute the above value of $x$ in $dx + ds + dt$, putting $ds = de + dz$, and we have $dt = (de + dz) + \frac{2g}{h}(z - x)$. To find $de$ in terms of $dz$, differentiate the equation of the curve, which in this position is $$x = r\operatorname{ver.}\sin^{-1}(z + r) + \sqrt{3}r - z^2;$$ whence $$ds = \frac{rdz}{\sqrt{3}r - z^2} = \frac{(r - z)dz}{\sqrt{3}r - z^2} = \frac{(2r - z)dz}{\sqrt{3}r - z^2};$$ $$\therefore de = \left[\frac{2r - 1}{q}\right]dz$$ ($r$=radius of the generating circle). Substituting, we have $$dt = \sqrt{\frac{r}{q}}\left[\frac{-de}{\sqrt{3}r - z^2}\right];$$ $$\therefore t = \sqrt{\frac{r}{q}}\int_{0}^{o}\frac{dz}{\sqrt{3}r - z^2} = \sqrt{\frac{r}{q}}\operatorname{ver.}\sin^{-1}\frac{x}{4h} = \pi\sqrt{\frac{r}{q}}.$$ Hence the whole oscillation occupies a time $= \pi\sqrt{\frac{r}{q}}$ (since $l = 4h$). This is independent of $h$, i.e., the oscillations are *isochronal*. This might have been proved by showing that $p_t$ is proportional to $OB$ measured along the curve; i.e., that the motion is *harmonic*. (§ 59, Prob. 2.) 78. Simple Circular Pendulum.—If the material point oscillates in a plane of a circle, Fig. 88, proceeding as in the preceding problem, we have finally, after integration by series, as the time of a full oscillation, $$t = \pi\sqrt{\frac{h}{g}}[1 + \frac{1}{8}\cdot\frac{h}{7} + \frac{9}{256}\cdot\frac{h^3}{7^3} + \cdots].$$ Hence for a small $h$ the time is nearly $\pi\sqrt{\frac{h}{g}}$, and the os- 89 MECHANICS OF ENGINEERING. Circulations nearly isochronal. (For the Compound Pendulum, see § 117.) 79. Change in the Velocity Square.—From eq. (7), § 74, we have $v^2 - c^2 = fpd$. But, from similar triangles, $du$ being the projection of any $de$ of the path upon the resultant force $R$ at that instant, $Rdu = Tds$ (or, Prin. of Virt. Veis. § 62, $Rdu = Tds + N \times 0$). $T$ and $N$ are the tangential and normal components of $R$. Fig. 80. Hence, finally, $$\frac{1}{2}Mv^2 - \frac{1}{2}M_0^2 = Rdu,$$ (a) for all elements of the curve between any two points. In general $R$ is different in amount and direction for each $de$ of the path, but $du$ is the distance through which $R$ acts, in its own direction, during the time that $de$ occurs; $Rdu$ is called the work done by $R$ when $de$ is described by the body. The above equation is read : The difference between the initial and final kinetic energy of a body = the work done by the resultant force in that portion of the path. (These phrases will be further spoken of in Chap. VII.) Application of equation (a) to a planet in its orbit about the sun.—Fig. 90. Here the only force at any instant is the attraction of the sun $C = C' - w'$ (see Prob. § 39), where $C'$ is a constant vector. As $u$ diminishes, $v$ increases, therefore $de$ and $du$ have contrary signs; hence equation (a) gives (c being the velocity at some initial point $O$) $$\frac{1}{2}Mv^2 - \frac{1}{2}M_0^2 = - C \int_{u_0}^{u_0'} du = C \left[ \frac{1}{u_0'} - \frac{1}{u_0} \right];$$ (b) :. $v_0' = \sqrt{c^2 + \frac{2C}{u_0'} \left( \frac{1}{u_0'} - \frac{1}{u_0} \right)}$, which is independent of the direction of the initial velocity $c$. Note: If $u_0 = \infty$, the last member of equation (b) would reduce to $C + u_0$, and is numerically the quantity called potential in the theory of electricity. Fig. 80. Fig. 90. CURVILINEAR MOTION OF A MATERIAL POINT. 83 Application of $q$, (a) to a projectile in vacuo.—$G$, the body's weight, is the only force acting, and $O$ therefore = $R$, while $M = G + g$. Therefore equation (a) gives $$\frac{Q}{g} \cdot v^2 - \frac{dy}{dx} = G \int_0^{y_0} dy = Gy_0$$ :. $v = \sqrt{v^2 + 2gy}$, which is independent of the angle $\alpha$, of projection. Application of equation (a) to a body sliding, without friction, on a fixed curved guide in a vertical plane; initial velocity $= c$ at $O$. Since there is some pressure at each point between the body and the guide, to consider the body free in space, we must consider the guide removed and that the body describes the given curve as a result of the action of the two forces, its weight $G$, and the pressure $P$ normal to the guide. Thus $G$ is constant, while $P$ varies from point to point, though always (since there is no friction) normal to curve. At any point, $R$ being the resultant of $G$ and $P$, project ds upon $R$, thus obtaining du ; on $G$, thus obtaining dy ; on $P$, thus obtaining zero. But by the principle of virtual velocities (see § 66) we have $dRds = Gdy + P \times zero = Gdy$, which substituted in eq. (a) gives $$\frac{1}{g} \left( v^2 - c^2 \right) = \int_0^{y_0} Gdy = G \int_0^{y_0} dy = Gy_0$$ :. $v = \sqrt{v^2 + 2gy}$, and therefore depends only on the vertical distance fallen through and the initial velocity, i.e., is independent of the form of the guide. As to the value of $P$, the mutual pressure between the guide and body at any point, since $XY$ must equal $Mo^2 + r$, $r$ being the variable radius of curvature, we have, as in § 77, $$P - G \sin \varphi = Mo^2 + r ; : . . P = G(\sin \varphi + (\varphi + gr))$$ As in general, $\varphi$ and $r$ are different from point to point of point 84 MECHANICS OF ENGINEERING. the path, $P$ is not constant. (The student will explain how $P$ may be negative on parts of the curve, and the meaning of this circumstance.) 80. Projectiles in Vacuo.—A ball is projected into the air (whose resistance is neglected, hence the phrase in vacuo) at an angle $\alpha$, with the horizontal ; required its path; assuming it confined to a vertical plane. Resolve the motion into independent horizontal ($X$) and vertical ($Y$) components. The force acting, being correspondingly replaced by its horizontal component = zero, and its vertical component $-g$. Similarly the initial velocity along $X = c_x = c \cos \alpha$, along $Y = c_y = c \sin \alpha$, along $X$ acceleration $= p_x = 0 + M = 0$, i.e., the $X$ motion is uniform, the velocity $v_x$ remains $= c_x = c \cos \alpha$, at all points, hence, reckoning the time from $O$, at the end of any time $t$ we have $$x = c(t \cos \alpha) \dots \dots (1)$$ In the $Y$ motion, $p_y = (-g) + M = -g$, i.e., it is uniformly retarded, the initial velocity being $c_y = c \sin \alpha$; hence, after any time $t$, the $Y$ velocity will be (see § 56) $v_y = c_y - gt$, while the distance $$y = c(t \sin \alpha) - \frac{gt^2}{2} \dots \dots (2)$$ Between (1) and (2) we may eliminate $t$, and obtain as the equation of the trajectory or path $$y = x \tan \alpha - \frac{gt^2}{2c} \cot \alpha$$ For brevity put $c' = 2gh$, $\lambda$ being the ideal height due to the velocity $c$, i.e., $c' + 2g$ (see § 53 ; if the ball were directed vertically upward, a height $\lambda = c' + 2g$ would be actually attained, $\alpha$, being $= 90^\circ$), and we have $$y = x \tan \alpha - \frac{g^2}{4h} \cot \alpha$$ This is easily shown to be the equation of a parabola, with its axis vertical. CURVILINEAR MOTION OF A MATERIAL POINT. 85 The horizontal range.—Fig. 94. Putting $y = 0$ in equation (3), we obtain $$x \left[ \tan \alpha_{*} - \frac{x}{4A \cos^2 \alpha_{*}} \right] = 0,$$ which is satisfied both by $x = 0$ (i.e., at the origin), and by $x = 4A \cos \alpha_{*} \sin \alpha_{*}$. Hence the horizontal range for a given $\alpha$ and $\alpha_{*}$ is $$x_{\text{max}} = 4A \cos \alpha_{*} \sin \alpha_{*} = 2A \sin 2\alpha_{*}.$$ For $y = y_{\text{max}}$, that is a maximum ($c$ remaining the same), being then $y = 2y_{\text{max}}$. Also, $\sin 2\alpha_{*} = \sin (180^\circ - 2\alpha_{*}) = \sin 2(90^\circ - \alpha_{*})$, therefore any two complementary angles of projection give the same horizontal range. Greatest height of ascent; that is, the value of $y$ maximum, $$y_{\text{max}}.$$—Fig. 94. Differentiate (3), obtaining $$\frac{dy}{dx} = \tan \alpha_{*} - \frac{x}{4A \cos^2 \alpha_{*}},$$ which, put $x = 0$, gives $x = 2A \sin \alpha_{*} \cos \alpha_{*}$ and this value of $x$ in (3) gives $y_{\text{max}} = A \sin^2 \alpha_{*}$. (Let the student obtain this more simply by considering the $F$ motion separately.) To strike a given point; $c$ being given and $\alpha_{*}$ required.—Let $x'$ and $y'$ be the co-ordinates of the given point, and $\alpha'_{*}$ the unknown angle of projection. Substitute these in equation (3), $k$ being known $= c + 2g$, and we have $$y' = x' \tan \alpha'_{*} - \frac{x'^2}{4A \cos^2 \alpha'_{*}}.$$ Put $\cos^2 \alpha'_{*} = \frac{1}{1 + x'^2},$$ and solve for $\tan \alpha'_{*}$, whence $$\tan \alpha'_{*} = [2h + x'^2 - x'^2 - 4hy'] / x'. . . . (4)$$ Evidently, if the quantity under the radical in (4) is negative, $\tan \alpha'_{*}$ is imaginary, i.e., the given point is out of range with the given velocity of projection $c = vgh$; if positive, $\tan \alpha'_{*}$ has two values, i.e., two trajectories may be passed through the point; while if it is zero, $\tan \alpha'_{*}$ has but one value. The envelope, for all possible trajectories having the same 86 MECHANICS OF ENGINEERING. initial velocity $c$ (and hence the same $b$); i.e., the curve tangent to them all, has but one point of contact with any one of them; hence each point of the envelope, Fig. 95, must have co-ordinates satisfying the condition, $4h - x^2 - 4hy = 0$; i.e., (see equation (4)), that there is but one trajectory belonging to it. Hence, dropping primes, the equation of the envelope is $4h - x^2 - 4hy = 0$, as a new origin, a new horizontal axis $x$, and reckon $y'$ positive downwards; i.e., substitute $x = z'$ and $y = h - y'$. The equation now becomes $z'^2 = 4hy'$; evidently the equation of a parabola whose axis is vertical, whose vertex is at $O'$, and whose parameter is $4h$ double the maximum horizontal range. $O$ is therefore its focus. The range on an inclined plane.—Fig. 96. Let $OC$ be the trace of the inclined plane; its equation $y = k \cos \alpha x$ (see page 73), combined with the equation of the trajectory (eq. 3), will give the co-ordinates of their intersection $C$. That is, substitute $y = x \tan \beta$ in (3) and solve for $x$, which will be the abscissa $x_1$ of $C$. This gives $$x_1 = \frac{a}{4h + \cos^2 \beta}$$ where $\alpha$ is constant, and $\beta$ varies. That is, required the value of $\alpha$ which renders $OC$ a maximum. Differentiating (5) with respect to $\alpha$, putting this derivative equal to zero, we have $(4h + \cos^2 \beta) (\cos \alpha \cos (\alpha - \beta) - \sin \alpha \sin (\alpha - \beta)) = 0$; whence $\cos (\alpha - \beta) = 0$; i.e., $2\alpha - \beta = 90^\circ$, or, $\alpha = 45^\circ + \frac{\pi}{4}$, for a maximum range. By substitution this maximum becomes known. The velocity at any point of the path is $v = \sqrt{v_x^2 + v_y^2} =$ A diagram showing a parabolic trajectory with an inclined plane. The diagram includes labels for points O, O', C, and various angles. Fig. 95 A diagram showing the relationship between the variables in the problem. It includes labels for points O, O', C, and various angles. Fig. 96 CURVILINEAR MOTION OF A MATERIAL POINT. 87 $$\sqrt{v^2 - B_0 \sin \alpha} + g t^2$$ (see the first part of this § 80); while the time of passage from $O$ to any point whose abscissa is $x$ is $$t = x + c \cos \alpha;$$ obtained from equation (1). E.g., to reach the point $B$, Fig. 94, we put $x = z$, $44 \sin \alpha \cos \alpha$, and obtain $t = 2c \sin \alpha + g$. This will give the velocity at $B = \sqrt{v^2 - c}$. 81. Actual Path of Projectiles.—Small jets of water, so long as they remain unbroken, give close approximations to parabolic paths, as also any small dense object, e.g., a ball of metal, having a moderate initial velocity. The course of a cannon-ball, however, with a velocity of 1300 to 1400 feet per second is much affected by the resistance of the air, the descending branch of the curve being much steeper than the ascending; see Fig. 96a. The equation of this curve has not yet been determined, but only the expression for the slope (i.e., dy : dx) at any point. See Pogg. Barlett's Mechanics, § 151 (in which the body is a sphere having no motion of rotation). Swift rotation about an axis, as well as an unsymmetrical form with reference to the direction of motion, alters the trajectory still further, and may deviate it from a vertical plane. The presence of wind would occasion increased irregularity. See Johnson's Encyclopaedia, article "Gunnery." 82. Special Problem (imaginary; from Weibach's Mechanics. The equations are not homogeneous.)—Suppose a material point, mass $M$, to start from the point $O$, Fig. 97, with a velocity $= 9$ feet per second along the $X$ axis, being sub- jected thereafter to a constant attractive $X$ force, of value $X' = 19 M$, and to a variable $Y$ force increasing with the time (in seconds, reckoned from $O$), Fig. 96. Fig. 97. 88 MECHANICS OF ENGINEERING. viz., $Y = 8Mt$. Required the path, etc. For the $X$ motion we have $p_x = X + M = 19$, and hence $$\int_{x_0}^{x_0'} dx = \int_{t_0}^{t_0'} dt = 12 \int_{t_0}^{t_0'} dt; \text{ i.e., } v_x = 12t;$$ and $$\int_{x_0}^{x_0'} dx = \int_{t_0}^{t_0'} dt; \text{ i.e., } x = 19 \int_{t_0}^{t_0'} dt = 6x. \quad (1)$$ For the $Y$ motion $p_y = Y + M = 8t$, $\therefore \int_{y_0}^{y_0'} dy = 8 \int_{t_0}^{t_0'} dt;$ i.e., $v_y + 9 = 4t,$ and $$\int_{y_0}^{y_0'} dy = \int_{t_0}^{t_0'} dt;$$ $\therefore y = 4 \int_{t_0}^{t_0'} dt - 9 \int_{t_0}^{t_0'} dt, \text{ or } y = 4t - 9t. \quad (2)$ Eliminate $t$ between (1) and (2), and we have, as the equation of the path, $$y = \pm \frac{4}{3}\left(\frac{x}{3}\right)^{\frac{1}{2}} + 9\left(\frac{y}{3}\right)^{\frac{1}{2}}; \quad . . . . \quad (3)$$ which indicates a curve of the third order. The velocity at any point is (see § 74, eq. (1)) $$v = v_x^{\prime} + v_y^{\prime} = 4t + 9. \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)$$ The length of curve measured from $O$ will be (since $v = de/dt$) $$s = \int_{e_0}^{e_0'} ds = \int_{t_0}^{t_0'} vdt = 4\int_{t_0}^{t_0'} dt + 9\int_{t_0}^{t_0'} dt = 4t + 9l. \quad (5)$$ The slope, tan $\alpha$, at any point is $v_x + v_y = (4t - 9) + 12l,$ and $\therefore \frac{d}{dt}\tan\alpha = \frac{4t - 9}{12l}. \quad . . . . (6)$ The radius of curvature at any point (§ 74, eq. (6)), substituting $v_x = 12t$, also from (4) and (6), is $$r = v^{\prime} + \left[v_x^{\prime}\tan\alpha\right]^{-1} = \frac{1}{12}(4t + 9)^{-1}, \quad . . . (7)$$ and the normal acceleration is $a' = v^{\prime}$ (eq. (4), § 74), becomes from (4) and (7) $p_n = 12(t, per square second), a constant. Hence the centripetal or deviating force at any point, i.e., the CURVILINEAR MOTION OF A MATERIAL POINT. 89 $EN$ of the forces $X$ and $Y$, is the same at all points, and = $M'x + r = 12.M$ From equation (5) it is evident that the curve is symmetrical about the axis $X$. Negative values of $t$ and $s$ would apply to points on the dotted portion in Fig. 97, since the body may be considered as having started at any point whatever, so long as all the variables have their proper values for that point. (Let the student determine how the conditions of this motion could be approximated to experimentally.) 83. Relative and Absolute Velocities.—Fig. 98. Let $M$ be a material point having a uniform motion of velocity $v_x$ along a straight groove cut in the deck of a steamer, which itself has a uniform motion of translation, of velocity $v_y$, over the bed of a river. In one second $M$ advances a distance $v_x$, along the groove, while simultaneously has moved a distance $v_y = AB$ with the vessel. The absolute path of $M$ during the second is evidently $\overline{w}$ (the diagonal formed on $v_x$ and $v_y$), which may therefore be called the *absolute velocity* of the body (considering the bed of the river as fixed); while $v_x$ is its relative velocity, i.e., relative to the vessel. If the motion of the vessel is not one of translation, the construction still holds good for an instant of time, but then the absolute velocity of that point of the deck over which $M$ is passing at this instant, and $v_x$ is $M$'s velocity relatively to that point alone. Conversely, if $M$ be moving over the deck with a given absolute velocity $\overline{w} = v_x$, being that of the vessel, the relative velocity $v_y$ may be found by resolving $w$ into two components, one of which shall be $v_x$; the other will be $v_y$. If $\overline{w}$ is the absolute velocity and direction of the wind, the vane on the headland will be parallel to $\overline{MT}$, i.e., to $\overline{w}$; the relative velocity $\overline{w'}$ of the vane is rolling, and the mast-head therefore describing a sinusoidal path, the direction of the vane varies periodically. A diagram showing a material point \( M \) moving along a groove in a steamer's deck. The diagram includes labels for various velocities and directions: \( v_x \) for horizontal component, \( v_y \) for vertical component, \( w \) for absolute velocity, \( v_{\text{relative}} \) for relative velocity to vessel, and \( \overline{w'} \) for relative wind velocity. Fig. 98 90 MECHANICS OF ENGINEERING. Evidently the effect of the wind on the sails, if any, will depend on $v$, the relative, and not directly on $w$, the absolute, velocity. Similarly, if $w$ is the velocity of a jet of water, and $v$, that of a water-wheel channel, which the water is to enter without sudden deviation, or impact, the channel-partition should be made tangent to $v$, and not to $w$. Again, the aberration of light of the stars depends on the same cause as that of the velocity of the earth's surface (being practically equal to that of the centre); $w$ is the absolute direction and velocity of the light from a certain star. To see the star, a telescope must be directed along $MT$, i.e., parallel to $v$, the relative velocity; just as in the case of the moving vessel, the groove must have the direction $MT$; if the moving material point, having an absolute velocity $w$, is to pass down the groove without touching its sides. Since the velocity of light is 186,000 miles per second = $w_1$, and since $\frac{1}{w_1} = \frac{1}{w}$, then $\frac{1}{w} = w_1$. The angle of aberration $SMT$, Fig. 98, will not exceed 20 seconds of arc; while it is zero when $w$ and $v$ are parallel. Returning to the wind and sail-boat,* it will be seen from Fig. 98 that when $v_1 = w$ or even > $w$, it is still possible for $v_2$ to be of such an amount and direction as to give, on a sail properly placed, a small wind-pressure, having a small fore-and-aft component, which in the case of an ice-boat may exceed the small fore-and-aft resistance of such a craft, and thus $v_2$ will be still (if not greater) than $v_1$. Thus boats sometimes travel faster than the wind which drives it. This has often been proved experimentally on the Hudson River. * See § 571 for the mechanics of the sail-boat. MOMENT OF INERTIA. 91 CHAPTER IV. MOMENT OF INERTIA. [Norm.--For the propriety of this term and its use in Mechanics, see §§ 114, 916, and 229; for the present we deal only with the geometrical nature of these two kinds of quantity.] § 85. Plane Figures.--Just as in dealing with the centre of gravity of a plane figure (§ 23), we had occasion to sum the series $f dF$, $dF$ being the distance of any element of area, $dF$, from an axis; so in subsequent chapters it will be necessary to know the value of the series $f^2 dF$ for plane figures of various shapes referred to various axes. This summation $f^2 dF$ of the products arising from multiplying each elementary area of the figure by the square of its distance from an axis is called the moment of inertia of the plane figure referred to the axis in question, its symbol being $I$. If this axis is perpendicular to the plane of the figure, it may be named the circular mom. of inertia ($\S$ 90-93); if the axis lies in the plane, the rectangular mom. of inertia of inertia ($\S$ 90-93). Since the $I$ of a plane figure evidently consists of four dimensions of length, it may always be resolved into two factors, thus $I = FP$, in which $F =$ total area of the figure, while $k = \sqrt{1 + P}$, is called the radius of gyration, because if all the elements of area were situated at the same radial distance, $k$, from the axis, the moment of inertia would still be the same, viz., $$I = f^2 dF = k^2 dF = FP.$$ § 86. Rigid Bodies.--Similarly, in dealing with the rotary motion of a rigid body, we shall need the sum of the series $f^3 dM$, meaning the summation of the products arising from multiplying the mass $dM$ of each elementary volume $dV$ of a 92 MECHANICS OF ENGINEERING. rigid body by the square of its distance from a specified axis. This will be called the moment of inertia of the body with respect to the particular axis mentioned (often indicated by a subscript), and will be denoted by $I$. As before, it can often be conveniently written $M R$, in which $M$ is the whole mass, and $k$ its "radius of gyration" for the axis used, $k$ being $\sqrt{I + M}$. If the body is homogeneous, the heaviness, $y$, of all its particles will be the same, and we may write $$I = f y^2 dM = (y + g) f y^2 dV = (y + g) V \rho.$$ 87. If the body is a homogeneous plate of an infinitely small thickness $= r$, and of area $= A$, we have $I = (y + g) f y^2 dV = (y + g) r f y^2 dF$, i.e., $(y + g) \times$ thickness $\times$ mom. inertia of the plane figure. 88. Two Parallel Axes. Reduction Formula—Fig. 90. Let $Z$ and $Z'$ be two parallel axes. Then $I_x = f y^2 dM$, and $I_y = f y'^2 dM$. But $d$ being the distance between the axes, so that $a' - b' = d$, we have $\rho' = (x - a') + (y - b') = (x' + y') + d - 2ax - 2by$, and . . . $$I_x = f y^2 dM + d^2 dM - 2ax dM - 2by dM.$$ (1) But $f y^2 dM = I_x$, $fdM = M$, and from the theory of the centre of gravity (see § 23, eq. 1), knowing that $dM = ydV + g$, and . . . that $(fy'dV) + (g'dM)$ we have $fdM = Mg$; hence (1) becomes $$I_x = I_y + Mg(d - 2ax - 2by).$$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2) in which $a$ and $b$ are the x and y of the axis $Z_1 Z_2$ and refer to the centre of gravity of the body. If $Z$ is a gravity-axis (call it $g$), both $a$ and $b$ = 0, and (2) becomes $$I_x = I_y + Mg.$$ or $$k_x^2 = k_y^2 + d^2.$$ . . . . . . . . . . . . . . . . . . . . . . . (3) It is therefore evident that the mom. of inertia about a gravity-axis is smaller than about any other parallel axis. Eq. (3) includes the particular case of a plane figure, by Fig. 90. MOMENT OF INERTIA. 93 writing area instead of mass, i.e., when $Z$ (now $\rho$) is a gravity-axis, $$I_x = I_y + Fd^2 \qquad \ldots \qquad \ldots \qquad (4)$$ 89. Other Reduction Formulae; for Plane Figures.—(The axes here mentioned lie in the plane of the figure.) For two sets of rectangular axes, having the same origin, the following holds good. Fig. 100. Since $$I_x = f y'dF, \quad \text{and} \quad I_y = f x'dF,$$ we have $$I_x + I_y = f(x^2 + y^2)dF.$$ Similarly, $$I_x + I_y = f(x^2 + y^2)dF.$$ But since the $x$ and $y$ of any $dF$ have the same hypotenuse as the $x$ and $y$, we have $x^2 + y^2 = x^2 + y^2$; i.e., $I_x + I_y = I_x + I_y$. Fig. 100. Let $X$ be an axis of symmetry; then, given $I_x$ and $I_y$ ($O$ is anywhere on $X$), required $I_z$. $U$ being an axis through $O$ and making any angle $\alpha$ with $X$. $$I_z = f(y'dF - x'x\sin\alpha)dF; \quad \text{i.e.,}$$ $$I_z = \cos\alpha f(y'dF - 2\sin\alpha\cos\alpha)dF.$$ But since the area is symmetrical about $X$, in summing up the product $xyzdF$, for every term at $-y'dF$, there is also a term at $-y'dF$ to cancel it; which gives $f(y'dF = 0$. Hence $$I_z = \cos\alpha I_x + \sin\alpha I_y.$$ The student may easily prove that if two distances $a$ and $b$ be set off from $O$ on $X$ and $Y$ respectively, made inversely proportional to $vT_x$ and $vT_y$, and an ellipse described on $a$ and $b$ as semi-axes; then the moments of inertia of the figure about Fig. 100a. 94 MECHANICS OF ENGINEERING. any axes through $O$ are inversely proportional to the squares of the corresponding semi-diameters of this ellipse; called therefore the Ellipse of Inertia. It follows therefore that the moments of inertia about all gravity-axes of a circle, or a regular polygon, are equal; since their ellipse of inertia must be a circle. Even if the plane figure is not symmetrical, an "ellipse of inertia" can be located at any point, and has the properties already mentioned; its axes are called the principal axes for that point. 90. The Rectangle.--First, about its base. Fig. 101. Since all points of a strip parallel to the base have the same co-ordinate, $s$, we may take $ds$ the area of such a strip for $dF = bds$; $$\therefore I_x = \int_{-b}^{b} s^2 ds = \frac{b^3}{3}$$ Secondly, about a gravity-axis parallel to base. $$dF = bds : I_y = f^2 dF = b \int_{-a}^{a} s^2 ds = \frac{a^3}{3}$$ Thirdly, about any other axis in its plane. Use the results already obtained in connection with the reduction-formulae of §§ 88, 89. 90a. The Triangle.--First, about an axis through the vertex and parallel to the base; i.e., $I_x$ in Fig. 103. Here the length of the side is variable; call it $y$. From similar triangles, $$y = (b + h)$$ $$\therefore I_x = \int_{-h}^{h} y^2 ds = (b + h) \int_{-h}^{h} s^2 ds = \frac{h^3}{3}$$ Secondly, about $g$, a gravity-axis parallel to the base. Fig. 104. From § 88, eq. (4), we have, since $F = \frac{bh}{2}$ and $$d = \frac{bh}{2}, I_y = I_x - F d = \frac{bh}{2} - \frac{bh}{2}. \frac{bh}{2} = \frac{bh^3}{6}$$ **MOMENT OF INERTIA.** Thirdly, Fig. 104, about the base ; $I_{B} = \frac{1}{3}$ From § 88, eq. (4), $I_{B} = I_{A} + Fd^2$, with $d = \frac{1}{4}h$; hence $I_{B} = \frac{1}{3}bh^2 + \frac{1}{4}bh^2 = \frac{7}{12}bh^2$. **91. The Circle.**—About any diameter, as in Fig. 105. Polar co-ordinates, $I_{A} = f'd^2F$. Here we take $dF = \text{area of an elementary rectangle} = \rho d\varphi.d\rho$, while $s = \rho\sin\varphi$. A circle with a point labeled 'O' at its center, and radii labeled 'a', 'b', 'c', 'd'. A line segment from 'O' to 'a' is drawn, forming an angle with the horizontal axis. Another line segment from 'O' to 'b' is drawn, forming an angle with the vertical axis. A third line segment from 'O' to 'c' is drawn, forming an angle with the horizontal axis. A fourth line segment from 'O' to 'd' is drawn, forming an angle with the vertical axis. Fig. 104 A circle with a point labeled 'O' at its center, and radii labeled 'a', 'b', 'c', 'd'. A line segment from 'O' to 'a' is drawn, forming an angle with the horizontal axis. Another line segment from 'O' to 'b' is drawn, forming an angle with the vertical axis. A third line segment from 'O' to 'c' is drawn, forming an angle with the horizontal axis. A fourth line segment from 'O' to 'd' is drawn, forming an angle with the vertical axis. Fig. 105 $I_{A} = \int (\rho\sin\varphi)^*pd\rho d\varphi = \int_0^\pi [\sin^2\varphi d\varphi]^{*}\rho d\rho$ $= \frac{\pi}{4}\int_0^\pi \sin^2\varphi d\varphi = \frac{\pi}{4}\int_0^\pi (1 - \cos2\varphi)d\varphi$ $= \frac{\pi}{4}\int_0^\pi [1 - \frac{1}{2}\cos2\varphi]d\varphi$ $= \frac{\pi}{4}\left[\frac{\pi}{2} - \frac{1}{4}\sin(\pi)\right] = \frac{\pi}{4}\left[\frac{\pi}{2} - 0\right].$ i.e., $I_{A} = \frac{1}{4}\pi r^2$. **92. Compound Plane Figures.**—Since $I = f'd^2F$ is an infinite series, it may be considered as made up of separate groups or subordinate series, combined by algebraic addition, corresponding to the separate components of a plane figure. In each component figure, such subordinate series being the moment of inertia of one of these component figures; but these separate moments must all be referred to the same axis. It is con- venient to remember that the (rectangular) $I$ of a plane figure remains unchanged if we conceive some or all of its elements shifted any distance parallel to the axis of refer- ence. E.g., in Fig. 106, the sum of the $I_{A}$ of the rectangle $CEH$ and that of $FD$ is equal to the $I_{B}$ of the imaginary rectangle 96 MECHANICS OF ENGINEERING. formed by shifting one of them parallel to $B$, until it touches the other; i.e., $I_x$ of $CE + I_y$ of $FD = \frac{bh^*}{3} (890)$. Hence the $I_y$ of the $\mathbb{T}$ shape in Fig. 106 will be $I_y$ of rectangle $AD - I_x$ of rect. $CE - I_x$ of rect. $FD$. That is, $$I_y = \frac{bh^*}{3} (b^*h^* - b_h^*) \dots (890), \dots (1)$$ About the gravity-axis, g, Fig. 106. To find the distance $d$ from the base to the centre of gravity, we may make use of eq. (3) of § 23, writing areas instead, of volumes, or, experimentally, having cut the given shape out of sheet-metal or card-board, we may balance it on a knife-edge. Supposing $d$ to be known by some such method, we have, from eq. (4) of § 88, since the area $F = bh - b_h$, $I_y = I_x - F d^2$; i.e., $I_y = \frac{bh^*}{3} (b^*h^* - b_h^*) \dots (890)$ The double-$\mathbb{T}$ (or $\mathbb{X}$), and the box forms of Fig. 106a, if symmetrical about the gravity-axis $g$, have moments of inertia alike in form. Here the gravity-axis (parallel to base) of each of the two component rectangles, of dimensions $b$ and $A$, $h$, and $h_*$ respectively. Hence by algebraic addition we have (§ 90), for either compound figure, $$I_y = \frac{bh^*}{3} (b^*h^* - b_h^*) \dots (890) \dots (3)$$ (If there is no axis of symmetry parallel to the base we must proceed as in dealing with the $\mathbb{T}$-form.) Similarly for the ring, $$I_y = \frac{bh^*}{3} (r^* - r_*^*) \dots (891) \dots (4)$$ Fig. 107, or space between two concentric circumferences, we have, about any diameter or g (§ 91). A diagram showing a circle with concentric circles around it. A diagram showing a circle with concentric circles around it. Fig. 107. Fig. 108. **MOMENT OF INERTIA.** 97 The rhombus about a gravity-axis, $g$, perpendicular to a diagonal, Fig. 108.—This axis divides the figure into two equal triangles, symmetrically placed, hence the $I_g$ of the rhombus equals double the moment of inertia of one triangle about its base; hence (§ 368) $$I_g = 2 \cdot \frac{1}{4} M(h)^2 = \frac{1}{2} Mh^2.$$ .......................... (5) (The result is the same, if either vertex, or both, be shifted any distance parallel to $AB$.) For practice, the student may derive results for the trapezoid; for the forms in Fig. 106, when the inner corners are rounded into equal quadrants of circles; for the double-$T$, when the lower flanges are shorter than the upper; for the regular polygon. If the plane figure be bounded, wholly or partially, by curves, it may be subdivided into an infinite number of strips, and the moments of inertia of these (referred to the desired axis) added by integration, if the equations of the curve are known; if not, Simpson's Rule, for a finite even number of strips, of equal width, may be employed for an approximate result. If these strips are parallel to the axis, the $I$ of any one strip = its length $\times$ its width $\times$ square of distances from axis; while if they are not parallel to the axis considered in, the axis, its $I = \frac{1}{3}$ its width $\times$ cube of its length (§ 368). A graphic method of determining the moment of inertia of any irregular figure will be given in a subsequent chapter. **94. Polar Moment of Inertia of Plane Figures (§ 85).—Since the axis is now perpendicular to the plane of the figure, intersecting it in point $O$, the distances of the elements of area will all radiate from this point, and would better be denoted by $\rho$ instead of $x$; hence, Fig. 109, $\rho^2 dF$ is the polar moment of inertia of any plane figure about a specified point $O$; this may be denoted by $I_\rho$. But $\rho^2 = x^2 + y^2$, for each $dF$; hence $$I_\rho = f(x^2 + y^2)dF = f'dF + g'dF = I_x + I_y.$$ 7 Fig. 109. 98 MECHANICS OF ENGINEERING. i.e., the polar moment of inertia about any given point in the plane equals the sum of the rectangular moments of inertia about any two axes of the plane figure, which intersect at right angles in the given point. We have therefore for the circle about its centre $$I_p = 4\pi r^2 + \frac{1}{2}\pi r^2 = \frac{3}{2}\pi r^2;$$ For a ring of radius $r$, and $r_c$, $$I_p = \frac{1}{2}(r^2 - r_c^2);$$ For the rectangle about its centre, $$I_p = \frac{1}{2}b^2h^2 + \frac{1}{2}a^2h^2 = \frac{1}{2}bh(b^2 + h^2);$$ For the square, this reduces to $$I_p = b^3.$$ (See §§90 and 91.) 95. Slender, Prismatic, Homogeneous Rod.—Returning to the moment of inertia of rigid bodies, or solids, we begin with that of a material line, as it might be called, about an axis through its extremity making some angle $\alpha$ with the rod. Let $l$ = length of the rod, $F$ its cross-section (very small, the result being strictly true only when $F = 0$). Subdivide the rod into an infinite number of small prisms, each having $F$ as its area. The mass of such a prism is $ydy$, where $y$ is the heaviness of the material; then the mass of an elementary prism, or $dM = (y + g)Fd$, while its distance from the axis Z is $\rho = \sin \alpha$. Hence the moment of inertia of the rod with respect to Z as an axis is $$I_z = \int y^2 dM = (y + g)\int Fd(y + g)\sin \alpha d\alpha.$$ But $yF\sin \alpha = \text{mass of rod and } l\sin \alpha = a$, the distance of the further extremity from the axis; hence $I_z = \frac{1}{2}Ma^2$ and the radius of gyration, or $k$, is found by writing $\frac{1}{2}Ma^2 = M\cdot k^2$; :. $k^2 = \frac{1}{4}a^2$, or $k = \sqrt{\frac{1}{4}}a$ (see § 86). If $\alpha = 90^\circ$, $a = l$. 96. Thin Plates. Axis in the Plate.—Let the plates be homogeneous and of small constant thickness $= r$. If the surface of **MOMENT OF INERTIA.** 90 the plate be $F$ and its heaviness $y$, then its mass is $yFr + g$. From § 87 we have for the plate, about any axis, $$I = (y + g)r \cdot mom. of inertia of the plane figure formed by the shape of the plate.$$ (1) Rectangular plate. Gravity-axis parallel to base.--Dimensions $b$ and $a$. From eq. (1) and § 90 we have $$L = (y + g)r \cdot \frac{1}{2}bh^2 = (yh + g)\frac{1}{2}bh^2 = \frac{1}{2}Ma; \therefore k' = \frac{1}{2}h^2.$$ Similarly, if the base is the axis, $I_b = \frac{1}{3}Ma; \therefore k' = \frac{1}{3}h^2$. Triangular plate. Axis through vertex parallel to base.--From eq. (1) and § 90, dimensions being $b$ and $h$, $$I_v = (y + g)\frac{1}{3}bh^2 = (yh + g)\frac{1}{3}bh^2 = \frac{1}{3}Ma; \therefore k' = \frac{1}{3}h^2.$$ Circular plate, with any diameter as axis.--From eq. (1) and § 91 we have $$I_p = (y + g)\pi r^2 = (\pi r^2 + g)\pi r^2 = \pi Mr^2; \therefore k' = \pi r^2.$$ 97. Plates or Right Prism of Any Thickness (or Altitude). Axis Perpendicular to Surface (or Base).--As before, the solid is homogeneous, i.e., of constant heaviness $y$; let the altitude be $h$. Consider an elementary prism, Fig. 11, whose length is parallel to the axis of reference $Z$. Its altitude is $A$ that of the whole solid; its base $= dF$ is an element of $F$ the area of the base of solid; and each point of it has the same $\rho$. Hence we may take its mass, $yhdF + g$, as the $dM$ in summing the series $$I_a = f yhdM;$$ $$I_x = f xhdM/dF;$$ $$I_y = (yh + g) \cdot polar mom. of inertia of base. . . .$$ (2) By the use of eq. (2) and the results in § 94 we obtain the following: Circular plate, or right circular cylinder, about the geometrical axis. $r =$ radius, $h =$ altitude. $$I_c = (yh + g)\pi r^2 = (yh + g)\pi r^2 = Ma; \therefore k' = \pi r^2.$$ Right parallelogramed or rectangular plate.--Fig. 113. $$I_b = (yh + g)\frac{1}{4}M(h_1 + b) = Ma; \therefore k' = h_1b.$$ 702834 A 100 MECHANICS OF ENGINEERING. For a hollow cylinder, about its geometric axis, $$I_x = (p + p) \pi (r_1^2 - r_2^2) = \frac{1}{2} M (r_1^2 + r_2^2) \cdot k' = \frac{1}{2} (r_1^2 + r_2^2).$$ Fig. 113. 88. Circular Wire.—Fig. 113 (perspective). Let $Z$ be a gravity-axis perpendicular to the plane of the wire; $X$ and $Y$ lie in this plane, intersecting at right angles in the centre $O$. The wire is homogeneous and of constant (small) cross-section. Since, referred to $Z$, each $dM$ has the same $\rho = p$, we have $$I_x = \frac{1}{2} p dM = Mr^2.$$ Now $I_x$ must equal $I_y$, and (§ 94) their sum is $I_z$. $$\therefore I_x,$ or $I_y = \frac{1}{2} Mp^2,$ and $$k_x,$ or $k_y = \frac{1}{2} p.$$ 90. Homogeneous Solid Cylinder, about a diameter of its base. —Fig. 114. Divide the cylinder into an infinite number of laminae, or thin plates, parallel to the base. Each is some distance $s$ from $X$, of thickness $dh$, and of radius $r$ (constant). In each draw a gravity-axis (of its own) parallel to $XY$. Add all these laminae to the $I_x$ of the whole cylinder by adding the $I_y$ of all the laminae. Thus, if any one lamina ($§ 96$, circular plate) is its mass $\times r^2$, hence its $$I_x (\text{eq. (3), § 88}) = \text{its } I_x + \text{(its mass)} \times s^2.$$ Hence for the whole cylinder $$I_x = \int_{-h}^{h} (\pi r^2 dz \times r^2 + g(4r^2 + s^2)) ds = (\pi r^4 + g(4r^2 + s^2)) \int_{-h}^{h} dz + (\pi r^4 + g(4r^2 + s^2)) ds;$$ i.e., $$I_x = (\pi r^4 + g)(4r^2 + s^2) = M(4r^2 + s^2).$$ 100. Let the student prove (1) that if Fig. 114 represent any right prism, and $k_p$ denote the radius of gyration of any one lamina, referred to its gravity-axis parallel to $X$, then the $I_x$ of whole prism is $$M(k_p^2 + \frac{1}{4}k_p^2);$$ and (2) that the moment **MOMENT OF INERTIA.** 101 of inertia of the cylinder about a gravity-axis parallel to the base is $I = M(r^2 + \frac{h^2}{3})$. 101. Homogeneous Right Cone.—Fig. 115. *First*, about an axis $V$, through the vertex and parallel to the base. As before, divide into laminae parallel to the base. Each is a circular thin plate, but its radius, $x$, is not $r$, but, from proportion, is $x = (r + h)z$. The $I$ of any lamina referred to its own gravity-axis parallel to $V$ is $(\S 96)$ = (its mass) $\times x^2$, and its $I_x$ (eq. (3), §88) is $= z = \text{its mass} \times z^2 + \frac{F_{xx}}{20} \times \text{its mass} \times z^2$. Hence for the whole cone, $I_x = \int (\pi x^2 dz + g)(x^2 + z^2)$ $= \frac{\pi r^2(1 - \frac{r}{h} + 1)}{4} \int_0^h dz = \frac{3}{20}(r^2 + 4h^2).$ *Secondly*, about a gravity-axis parallel to the base.—From eq. (3), §88, with $d = \frac{h}{4}$ (see Prob. 7, §26), and the result just obtained, we have $I = Mg(\frac{r^2}{4} + \frac{h^2}{3}).$ *Thirdly*, about its geometric axis, Z.—Fig. 116. Since the axis is perpendicular to each circular lamina through the centre, its $I_z$ ($\S 97$) is $= \text{its mass} \times (\text{rad.})^2 = (\pi r^2dz + g)\times z^2.$ Now $x = (r + h)z$, and hence for the whole cone $I_z = \int (\pi r^2dz + gh)\int_0^h z^2 dz = (\pi r^2dz + g)\times r^2 = M_{\text{vol}}r^2.$ Fig. 115. Fig. 116. Fig. 117. 102. Homogeneous Right Pyramid of Rectangular Base.—About its geometrical axis. Proceeding as in the last para- 102 MECHANICS OF ENGINEERING. graph, we derive $I_x = M_{xy}d^2$, in which $d$ is the diagonal of the base. 103. Homogeneous Sphere.—About any diameter. Fig. 118. $I_x = I$ Divide into lamina perpendicular to $Z$. By § 97, and noting that $x^2 = r^2 - z^2$, we have finally, for the whole sphere, $$I_x = (y\pi + 2g)\left[\frac{r^2}{2} + \frac{r^2}{2}\left(r^2 - \frac{z^2}{r^2} + \frac{z^2}{r^2}\right) = \frac{r^2}{2}\pi r^3 + g\right.$$ $$= (\frac{1}{2}\pi r^3 + g)r^3 = M_{xy}r^3 : k_x = \frac{1}{4}\pi r^3.$$ For a segment, of one or two bases, put proper limits for $s$ in the foregoing, instead of $+r$ and $-r$. 104. Other Cases.—Parabolic plate, Fig. 119, homogeneous and of (any) constant thickness, about an axis through $O$, the middle of the chord, and perpendicular to the plate. Fig. 119. This is Fig. 120. $$I = M_1(4e^a + 4b^a).$$ The area of the segment is $\frac{4ab}{\pi}$. For an elliptic plate, Fig. 120, homogeneous and of any constant thickness, semi-axes $a$ and $b$, we have about an axis through $O$, normal to surface $I_o = M_1(a^2 + b^2)$; while for a very small constant thickness $$I_x = M_1b^a,$$ and $$I_y = M_1a^a.$$ The area of the ellipse is $\pi ab$. Considering Figs. 119 and 120 as plane figures, let the student determine their polar and rectangular moments of inertia about various axes. (For still other cases, see p. 518 of Rankine's Applied Mechanics, and pp. 593 and 594 of Coxe's Weisbach.) 105. Numerical Substitution.—The moments of inertia of plane figures involve dimensions of length alone, and will be utilized in the problems involving flexure and torsion of beams, where the inch is the most convenient linear unit. E.g., the MOMENT OF INERTIA. 108 polar moment of inertia of a circle of two inches radius about its centre is $\frac{4}{3}\pi r^2 = 25.13$ square dimension, or four-dimension, inches, as it may be called. Since this quantity contains four dimensions on each, the use of the foot instead of the inch would diminish its numerical value in the ratio of the fourth power of twelve to unity. The moment of inertia of a rigid body, or solid, however, $$= M r^2 = (G + g) k^2,$$ in which $G$, the weight, is expressed in units of force, $g$ involves both time and space (length), while $k^2$ involves length (two dimensions). Hence in any homogeneous formula in which the $I$ of a solid occurs, we must be careful to employ units consistently; e.g., if in substituting $G + g$ for $M$ (as will always be done numerically) we put $g = 32\frac{1}{2}$, we should use the second as unit of time, and the foot as linear unit. 108. Example.—Required the moment of inertia, about the axis of rotation, of a pulley consisting of a rim, four parallelepipedal arms, and a cylindrical hub which may be considered solid, being filled by a portion of the shaft. Fig. 121. Call the weight of the hub $G$, its radius $r_1$; similarly, for the rim $G_r r$, and $r_2$; the weight of one arm being $= G_a$. The total $I$ will be the sum of the $I$'s of the component parts referred to the same axis. viz.: Those of the hub and rim will be $(G + g) r_1^2$ and $(G_r + g) r_2^2$, respectively ($§ 97), while if the arms are not very thick compared with their length, we have for them ($§§ 95 and 88) $$4(G + g)\left[(r_1 - r)^2 - (r_2 - r)^2\right] + [r + k(r_1 - r)]^2,$$ as an approximation (obtained by reduction from the axis at the extremity of an arm to a parallel gravity-axis, then to the required axis, then multiplying by four). In most fly-wheels, the rim is proportionally so heavy, besides being the farthest removed from the axis of rotation, that the moment of inertia of the other parts may be for practical purposes neglected. Fig. 121. 104 MECHANICS OF ENGINEERING. **107. Ellipsoid of Inertia.—The moments of inertia about all axes passing through any given point of any rigid body whatever may be shown to be inversely proportional to the squares of the diameters which they intercept in an imaginary ellipsoid, whose centre is the given point, and whose position in the body depends on the distribution of its mass and the location of the given point. The three axes which contain the three principal diameters of the ellipsoid are called the *Principal Axes* of the body for the given point. This is called the *ellipsoid of inertia.* (Compare § 89.) Hence the moments of inertia about these axes are equal, since then the ellipsoid becomes a sphere. It can also be proved that for any rigid body, if the co-ordinate axes $X$, $Y$, and $Z$ are taken coincident with the three principal axes at any point, we shall have $$f_{y}dM = 0; \quad f_{z}dM = 0; \quad \text{and} \quad f_{x}dM = 0.$$ DYNAMICS OF A RIGID BODY. 106 CHAPTER V. DYNAMICS OF A RIGID BODY. 108. General Method.--Among the possible motions of a rigid body the most important for practical purposes (and fortunately the most simple to treat) are: a motion of translation, in which the particles move in parallel right lines with accelerations and velocities at any given instant; and rotations about an axis, in which the particles move in circular parallel planes with velocities and accelerations proportional (at any given instant) to their distances from the axis. Other motions will be mentioned later. To determine relations, or equations, between the elements of the motion, the mass and form of the body, and the forces acting (which do not necessarily form an unbalanced system), the most direct method to be employed is that of two equivalent systems of forces ($\S$ 15), one consisting of the actual forces acting on the body, considered free, or imaginary, consisting of the infinite number of forces equal, but directed in opposite directions composing the body, would account for their individual actions, as if they were an assembly of particles without mutual actions or coherence. If the body were at rest, then considered free, and the forces referred to three co-ordinate axes, they would constitute a balanced system, for which the six summations $X_1 X_2 X_3 Z_1 Z_2 Z_3 (\text{mom})_x$, $(\text{mom})_y$, and $(\text{mom})_z$ would each equal zero; but in most cases of motion some or all of these sums are equal (not at any given instant), not to zero, but to the corresponding sum of moments of the second kind, i.e., to expressions involving the masses of the particles (or material points), their distribution in the body, and the 108 MECHANICS OF ENGINEERING. elements of the motion. That is, we obtain six equations by putting the $\mathbb{Z}X$ of the actual system equal to the $\mathbb{Z}X$ of the imaginary, and so on; for a definite instant of time (since some of the quantities may be variable). 109. Translation—Fig. 129. At a given instant all the particles have the same velocity $v$, in parallel right lines (parallel to the axis $X$, say), and the same acceleration $a$. Required the $\mathbb{Z}X$ of the acting forces, shown at (I). (II) shows the imaginary equivalent system, consisting of a force $=dMg$ applied parallel to $X$ to each particle, since such a force would be necessary (from eq. (IV.) §55) to account for the accelerated rectilinear motion of the particle, independently of the others. Putting $(\mathbb{Z}X)_t = (\mathbb{Z}X)_B$ we have $$(\mathbb{Z}X)_t = pdM = pg/dM = Mpg \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \\ (\mathbb{Z}X)_t = pdM = pg/dM = Mpg = -pg/dM = -p.dM = -pdM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM = -p.dM ... \\ It is evident that the resultant of system (II) must be parallel to $X$, hence that of (I), which is $(\mathbb{Z}X)$ and may be denoted by $R$, must also be parallel to $X$; let $a$ be perpendicular distance from $R$ to the plane $XY$: $a$ will be parallel to $Z$. Now put $(\mathbb{Z}(\text{mom.})_Y)_r=(\mathbb{Z}(\text{mom.})_Y)_B(Y)$ an axis perpendicular to paper through $O$ and we have $-Ra=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt}Mpz=-\frac{d}{dt} Mp z \\ Hence, if a rigid body has a motion of translation, the resultant force must act in a line through its centre of mass (hence through its centre of mass), and parallel to the direction of motion. Or, practically, in dealing with a rigid body having a motion of translation, we may consider it concentrated at its centre of mass. If the velocity of translation is uniform, $R=M\times0=0$, i.e., the forces are balanced. DYNAMICS OF A RIGID BODY. 107 110. Rotation about a Fixed Axis.—First, as to the elements of space and time involved. Fig. 123. Let $O$ be the axis of rotation (perpendicular to $Y$), $OY$ a fixed line of reference, and $OA$ a convenient line of the rotating body, passing through the axis and perpendicular to $OY$. The motion of the body in its angular motion, which is the same as that of $OA$. Just as in linear motion we dealt with linear space ($x$), linear velocity ($v$), and linear acceleration ($p$), so here we distinguish at any instant; $\alpha$, the angular space between $OY$ and $OA$; $\omega = \frac{d\alpha}{dt}$, the angular velocity, or rate at which $\alpha$ is changing; and $\theta = \frac{d\omega}{dt} = \frac{d^2\alpha}{dt^2}$, the angular acceleration, or rate at which $\omega$ is changing. These are all reckoned in $\pi$-measure and may be + or -, according to their direction against or with the hands of a watch. (Let the student interpret the following cases: (1) at a certain instant $\alpha$ is +, and $\theta = -$; (2) $\omega$ is -, and $\theta = +$; (3) $\alpha$ is -, $\omega$ and $\theta$ both +; (4) $\alpha +$, $\omega$ and $\theta$ both -.) For rotary motion we have therefore, in general, $$\omega = \frac{d\alpha}{dt}; \quad \theta = \frac{d\omega}{dt} = \frac{d^2\alpha}{dt^2}; \quad (\text{VII})$$ and $$\therefore \omega d\alpha = d\theta dt; \quad (\text{VIII})$$ corresponding to eqs. (I.), (II.), and (III.) in § 80, for rectilinear motion. Hence, for uniform rotary motion, $\omega$ being constant and $\theta = 0$, we have $\alpha = \omega t$, $t$ being reckoned from the instant when $\alpha = 0$. For uniformly accelerated rotary motion $\theta$ is constant, and 108 MECHANICS OF ENGINEERING. if $\omega$ denote the initial angular velocity (when $a$ and $t = 0$), we may derive, precisely as in § 56, $$\omega = \omega_0 + \theta t; \quad . . . \quad (1)$$ $$\alpha = \omega dt + \theta dt^2; \quad . . . \quad (2)$$ $$\alpha = \frac{d}{dt}(\omega dt); \quad . . . \quad (3)$$ and $\alpha = (\omega dt + \omega dt^2). \quad . . . \quad (4)$ If in any problem in rotary motion $\theta$, $\omega$, and $\alpha$ have been determined for any instant, the corresponding linear values for any point of the body whose radial distance from the axis is $r$, will be $v = ar (= distance described by the point measured along its circular path from its initial position), v = ar = its velocity, and $p_t = \theta r$ its tangential acceleration, at the instant in question. Examples.—(1) What value of $\omega$, the angular velocity, is implied in the statement that a pulley is revolving at the rate of 100 revolutions per minute? 100 revolutions per minute is at the rate of $\pi \times 100 = 628.32$ (measure units) of angular space per minute $= 10.473$ per second; i.e., $\omega = 628.32$ per minute or 10.473 per second. (2) A grindstone whose initial speed of rotation is 90 revolutions per minute is brought to rest in 30 seconds, the angular retardation (or negative angular acceleration) being constant; required the angular acceleration, $\theta$, and the angular space $s$ described. Use the second as unit of time. $$\omega_i = \frac{576}{30} = 9.424 \text{ per second}; \quad . . . \text{from eq. (1)}$$ $$\theta = -\frac{\omega_i}{30} = -9.424 / 30 = -0.3141 (\pi-\text{measure units})$$ per "square second." The angular space, from eq. (3), is $$s = \omega_i t + \theta t^2 = 9.424 - 4(0.3141)^2/30 = 14.13$$ ($\pi$-measure units), i.e., the stone has made 29.4 revolutions in coming to rest and a point 2 ft. from the axis has described a distance $s = ar = 14.13 \times 2 = 28.26$ ft. in its circular path. 111. Rotation. Preliminary Problem. Axis Fixed.—For clearness in subsequent matter we now consider the following DYNAMICS OF A RIGID BODY. 100 simple case. Fig. 124 shows a rigid body, consisting of a drum, an axle, a projecting arm, all of which are imponderable, and a single material point, whose weight is $G$ and mass $M$. An imponderable flexible rod, in which the tension $P_1$ keeps constant and $P_2$ unwinds from the drum. The axe coincides with the vertical axis $Z$, while the cord is always parallel to $Y$. Initially (i.e., when $t = 0$) $M$ lies at rest in the plane $ZY$. Required its position at the end of any time $t$ (i.e., at any instant) and also the reactions of the bearings at $O$ and $O'$, supposing no vertical pressure to exist at $O_1$ and that $F$ and $M$ are at the same level. No friction. At any instant the equilibrium condition $\Sigma X = 0$, $\Sigma Y = 0$, etc., may be found from the six equations formed by putting $\Sigma X$, etc., of the system of forces in Fig. 124, equal, respectively, to the $\Sigma X$, etc., of the imaginary equivalent system in Fig. 125, and two others to be mentioned subsequently. Since, at this instant, the velocity of $M$ must be $v = \omega x$, its tangential acceleration $p_t = \theta p$, its circular motion could be produced, considering it free (eq. (5), § 74), by a tangential force $T = \frac{d}{dt}(\frac{d}{dt}\omega x)$, i.e., $\frac{d}{dt}(\frac{d}{dt}\omega x) + \frac{d}{dt}(\frac{d}{dt}\omega x) = T$. Hence the system in Fig. 125 is equivalent to that of Fig. 124, and from putting the $\Sigma(mom.)_z$ of one = that of the other, we derive $$Pa = Tp; \text{i.e}, Pa = \theta Mp.$$ (1) whence $\theta$ becomes known, and is evidently constant, since $P$, $a$, $M$, and $\rho$ are such. . . . The angular motion is uniformly accelerated, and from eqs. (2) and (3) $\frac{d}{dt}\omega = \theta\omega$. . . . (3) Putting ($\Sigma Z$ of 124) = ($\Sigma Z$ of 125), gives $$Z - G = 0; \text{i.e}, Z = G.$$ (4) Fig. 124. Fig. 125. 110 MECHANICS OF ENGINEERING. Proceeding similarly with the $ZX$ of each system, $$X + X_1 = T \cos \alpha - N \sin \alpha = M P \cos \alpha - w^2 M P \sin \alpha, (5)$$ and with the $ZY$ of each, $$P + Y_1 + Y_2 = -T \sin \alpha - N \cos \alpha = -M P \sin \alpha$$ $$- w^2 M P \cos \alpha; (6)$$ while with the $ZX$ (mom.) we have, concieving all the forces in each system projected on the plane $ZY$ (see § 35), and noting that $y = P \cos \alpha$ and $x = P \sin \alpha$, $$+ w^2 P \cos \alpha + Y_1 + Y_2 = -(M P \sin \alpha) - (w^2 M P \cos \alpha) b, (7)$$ and with the $ZY$ (mom.), $$- w^2 P \sin \alpha - X_1 = -(M P \cos \alpha) + (w^2 M P \sin \alpha)b. . . (8)$$ From (7) we may find $Y_1$; from (8), $X_1$; then $X$, and $Y$, from (6) and (6). It will be noted that as the motion proceeds $\theta$ remains constant; $w$ increases with the time, $\alpha$ with the square of the time, and so on. For example, if $P = 4$, $T = 4$, $Y_1$, $X_1$, and $Y_1$ have variable values dependent on $\cos \alpha$ and $\rho \sin \alpha$, i.e., on the co-ordinates $y$ and $x$ of the moving material point. 112. Particular Supposition in the Preceding Problem with Numerical Substitution.—Suppose we have given (using the foot-pound-second system of units in which $g = 32.2$) $M = 64.4$ lbs., whence $$M = (G + g) = G = 4\) lbs.; \(t = 4\) ft.; \(s = 2\) ft.; \(v = 2\) ft.; and \(p = 4\) ft.; and that \(M\) is just passing through the plane \(ZX\), i.e., that \(a = 4\)\. We obtain, first, the angular acceleration, eq. (1), $$\theta = Pa + Mp' = 8 + 32 = 0.25 = 4.$$ From eqs. (2) and (3) we have at the instant mentioned (noting that when \(a\) was \(= 0\), \(t\) was \(= 0\)) $$w^2 a = 2a\theta = 4\pi = 0.7854 +,$$ while (2) gives, for the time of describing the quadrant, $$t = w + \theta = 3.544... seconds.$$ Since at this instant $\cos a = 0$ and $\sin a = 1$, we have, from (7), $$+0 + Y_1\times4 + 4\times2 - 4\times9\times4\times2; . . . Y_1 = -3\) lbs. DYNAMICS OF A RIGID BODY. 111 The minus sign shows it should point in a direction contrary to that in which it is drawn in Fig. 124. Eq. (6) gives $$-64.4 \times 4 - X_1 \times 4 = -0.4 \times 4 \times 5 \times 3 \times Z_1; Z_1 = -67.54 \text{ lbs}.$$ And similarly, knowing $X_1$ and $Z_1$, we have from (6) and (6), $$X_2 = +61.26 \text{ lbs}, \quad Y_2 = -3.00 \text{ lbs}.$$ The resultant of $X_1$ and $Y_1$, also that of $X_2$, $Y_2$, and $Z_2$, can now be found by the parallelogram (and parallelepiped) of forces, both in amount and position, noting carefully the directions of the components. These resultantts are the actions of the support on the body, and the action of the body against the support, at the instant considered (when $M$ is passing through the plane $ZT$; i.e., with $\alpha = \frac{\pi}{4}$). (At the same instant, suppose the string to break; what would be the effect on the eight quantities mentioned?) 113. Centre of Percussion of a Rod suspended from one End.—Fig. 126. The rod is initially at rest (see (I.) in figure), is straight homogeneous, and of constant (small) cross-section. Neglect its weight. A horizontal force or pressure, $P$, due to a blow (and varying with time), applied to the left end (as shown below), now acts upon it from the left, perpendicularly to the axis, $Z_0$ of suspension. An accelerated rotary motion begins about the fixed axis $Z$. (II.) shows the rod free, at a certain instant, with the reactions $X_0$ and $Y_0$ put in at $O_p$. (III.) shows an imaginary system which would produce the same effect at this instant, and consisting of a $dT = dM_{\theta0}$, and a couple $dM dZ$ applied to each $dM$, the rod being composed of an infinite number of such elements, all revolving about the axis. Considering that the rotation has just begun, $\omega$, the angular velocity is as yet small, and will be neglected. Required $Y_0$, the horizontal reaction of the support at $O$ in terms of $P$. By putting $\Sigma Y_{\theta0} = \Sigma Y_{\theta0}$, we have $$P - Y_0 = f dT = 8/4 dM = 8 M P.$$ 112 MECHANICS OF ENGINEERING. $\therefore Y_{z}=P-\theta M\bar{p};\ \bar{p}$ is the distance of the centre of gravity from the axis (N.B. $\theta M\bar{p}$ is only true when all the $\theta$'s are parallel to each other). But the value of the angular acceleration at this instant depends on $P$ and, for (mom.) $\bar{a}$ (in I.II), where $P=\mathscr{Z}(\mathrm{mom.})_{z}$ in (III), whereas $P=\frac{3}{4}\theta M^{2}L_{z}^{2}+M_{z}^{2}$, where $L_{z}$ is the moment of inertia of the rod about $Z$, and from § 95 $=M\bar{p}$. Now $\bar{p}=\frac{3}{4}\bar{a}$; hence, finally, $$Y_{z}=P\left[1-\frac{3}{4}\cdot\frac{a}{l}\right].$$ If now $Y_{z}$ is to be 0, i.e., if there is to be no shock between the rod and axis, we need only apply $P$ at a point whose distance $a=\frac{3}{4}l$ from the axis; only then $Y_{z}=0$. This point is called the centre of percussion for the given rod and axis. It and the point of suspension $O$ are interchangeable (see 118). (Lay a pencil on a table; tap it at a point distant one third of the length of the pencil from its end; it will begin to rotate about a horizontal axis through the farther end. Tap it at one end ; it will begin to rotate about a vertical axis through the point first mentioned. Such an axis of rotation is called an axis of instantaneous rotation, and is different for each point of impact—just as the point of contact of a wheel and rail is the one point of the wheel which is momentarily at rest, and about which, therefore, all the others are turning for the instant. Tap the pencil at its centre of gravity, and a motion of translation begins; see § 106.) 114. Rotation. Axis Fixed. General Formula.—Consider Fig. 127. Fig. 128. ing now a rigid body of any shape whatever, let Fig. 19T indicate the system of forces acting at any given instant, $Z$ being DYNAMICS OF A RIGID BODY. 118 the fixed axis of rotation, $a$ and $\theta$ the angular velocity and angular acceleration, at the given instant. $X$ and $Y$ are two axes, at right angles to each other and to $Z$, fixed in space. At this instant each $dM$ of the body has a definite $x$, $y$, and $\varphi$ (see Fig. 128), which will change, and also $a_x$ and $a_y$, which will not change, as the motion progresses, and is pursuing a circular path with a velocity $= \omega x$ and a tangential acceleration = $\theta \rho$. Hence, if to each $dM$ of the body (see Fig. 128) we imagine a tangential force $dT = dM \theta \rho$ and a normal force $-dM(\omega y) + \mu = \omega^2 dM \rho$ to be applied (eq. (5); § 74), and these alone act on the body, then the resultant of all these forces, all parallel to $XY$, etc., represent the sum (six) for Fig. 127. Let $\Sigma X$, etc., represent the sum (six) for Fig. 127, whatever they may be in any particular case, while for 128 we shall write the corresponding sums in detail. Noting that $$fdN\cos\varphi = \omega^2/dM\rho\cos\varphi = \omega^2/dMy = \omega^2My/(38);$$ that $fdN\sin\varphi = \omega^2/dM\rho\sin\varphi = \omega^2/dMz = \omega^2Mz;$$ and similarly, that $fdT\cos\varphi = 0/Mz;$$ while in the moment suma (the moment of $dT\cos\varphi$ about $Y$, for example, being $-dT\cos\varphi = -\theta dM\rho(\cos\varphi)z = -0dMyz$, the sum of the moments, of all the $(dT\cos\varphi)y = -0/dMyz)$. $$fdT\cos\varphi = 0/dMyz; fdN\sin\varphi = \omega^2/dMz,$$ we have, since the systems are equivalent, $$\Sigma X = +0My - \omega^2Mz; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (IX).$$ $$\Sigma Y = -0Mz - \omega^2Mg; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . (X).$$ $$\Sigma Z = 0; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; \quad ; (XI).$$ $\Sigma$ moma$_x$ = $-0/dMyz - \omega^2/dMyz; \quad ; (XII).$ $\Sigma$ moma$_y$ = $-0/dMyg + adfMz; \quad ; (XIII).$ $\Sigma$ moma$_z$ = $0/dMg = 0/Iz; \quad ; (XIV).$ These hold good for any instant. As the motion proceeds $x$ and $y$ change, as also the sums $fdMxz$ and $fdMyz$. If the body, however, is homogeneous, and symmetrical about the plane $XY$, $fdMxz$ and $fdMyz$ would always be zero; since 114 MECHANICS OF ENGINEERING. the $s$ of any $dM$ does not change, and for every term $dMy(+x)$, there would be a term $dMy(-x)$ to cancel it; similarly for $fdMz$. The eq. (XIV.), $\Sigma$ (moment about axis of rotat.) = $fdTp$, or $fdMp$ = (angular of inertia) × (sum of inertia of body about axis of rotat.), shows how the sum $fdMp$ arises in problems of this chapter. That a force $dT$ = $dMp$ should be necessary to account for the acceleration (tangential) $\theta p$ of the mass $dM$, is due to the so-called inertia of the mass ($\S 54$), and its moment $dTp$, or $dMp$, might, with some reason, be called that of the whole body. But custom has restricted the name to the sum $fdMp$, which, being without the $\theta$, has no term to suggest the idea of inertia. For want of a better name it is still retained, however, and is generally denoted by $I$. (See $\S\S 60$, etc.) 115. Example of the preceding. — A homogeneous right parallelopiped is mounted on a vertical axe (no friction), as in figure. $O$ is at its centre of gravity, hence both $x$ and $y$ are zero. Let its heaviness be $\gamma$, its dimensions $h$, $b_x$ and $b_y$ (see $\S 97$). $XY$ is a plane of symmetry, hence both $fdMx$ and $fdMz$ are zero at all times (see above). The tension $P$ in the (inextensible) cord is caused by the hanging weight $P_0$ (but not = $P_0$, unless the rotation is uniform). The figure shows both rigid bodies free from friction, i.e., undergoing translation; the parallelopiped, one of rotation about a fixed axis. No masses are considered except $P_0 + g$, and $bb_y + g$. The $I_z = Mgz$ of the latter is its mass $\times \frac{1}{2}(b_x^2+b_y^2)$, $\S 97$. At any instant, the cord being taut, if $p =$ linear acceleration of $P_0$, we have $$p = 0a.$$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . eq. $(a)$ From (XIV.), $Pa = 0I_z$ : : $P = 0I_z + a.$ . . . . . . . . . . . (1) Fig. 13. DYNAMICS OF A RIGID BODY. For the free mass $P + g$ we have (§ 109) $P - P =$ $$= (P + g) \theta a = P(1 - \theta a + g).$$ Equate these two values of $P$ and solve for $\theta$, whence $$\theta = \frac{P}{Mg + (P + g)a^2}.$$ (8) All the terms here are constant, hence $\theta$ is constant; therefore the rotary motion is uniformly accelerated, as also the translation of $P$. The formulae of § 56, and (1), (2), (3), and (4) of § 109, are applicable to this case, though $\theta$ is constant; see eq. (8). As for the five unknown reactions (components) at $O$, and $O_n$, the bearings, we shall find that they too are constant; for from (IX.) we have $$X_1 + X_2 = 0;$$ (4) from (X.) we have $$P + Y_1 + Y_2 = 0;$$ (5) from (XI.) we have $$Z_1 - Z_2 = 0;$$ (6) from (XII.) we have $$P \cdot AO + Y_1 \cdot O_1O - Y_2 \cdot O_2O = 0;$$ (7) from (XIII.) we have $$- X_1 \cdot O_1O + X_2 \cdot O_2O = 0.$$ (8) Numerical substitution in the above problem—Let the parallelogram be of wrought-iron; let $F = 48$ lbs.; $a = 6$ in. = $\frac{3}{4}$ ft.; $b = 3$ in. = $\frac{3}{4}$ ft. [see Fig. 119]; $h = 2$ ft. 3 in. = $\frac{5}{4}$ ft.; and $h = 4$ in. = $\frac{1}{4}$ ft. Also let $O_1O_2 = 18$ in. = $\frac{3}{4}$ ft., and $AO = 3$ in. = $\frac{3}{4}$ ft. Selecting the foot-pound-second system of units, in which $g = 32.2$, the linear dimensions must be used in lbs. per cubic foot, i.e., $y = 480$ [see § 8], and all forces in lbs., times in feet. The weight of the iron will be $G = Fy = h_bhy = \frac{3}{4} \times 480 = 90$ lbs.; its mass is $90 + 32.2g = 93.2$ ; its moment of inertia about $Z$ is $I_z = Mz^2 = My(h + b) = \frac{5}{7} \times 480 = 1.19$. (That is, the radius of gyration, $k_y$ is $\sqrt{\frac{480}{93.2}} = 0.635$ ft.; or the moment of inertia, or any result depending solely upon it, is just the same as if the mass were concentrated in a thin shell, or a line, or a point, at a distance of 0.635 feet from the axis.) We can now compute the angular acceleration, $\theta$, from eq. (8); 116 MECHANICS OF ENGINEERING. $$\theta = \frac{45 \times \frac{3}{8}}{24} = 1.191 + 0.375 = 1.566$$ $m$-measure units per "square second." The linear acceleration of $P$, $a_p = \theta u = 7.68$ feet per square second for the uniformly accelerated motion. Nothing has yet been said of the velocities and initial conditions of the motions; for what we have derived so far applies to any point of time. Suppose, then, that the angular velocity $\omega = 0$ when the time, $t = 0$; and correspondingly the velocity, $v = \omega t$, of translation of $P$, be also $= 0$ when $t = 0$. At the end of any time $t$, $\omega = \theta t$ (§§ 56 and 110) and $v = pt$. $\theta = 6at$; also the angular space, $\alpha = \theta t$, described by the parallotipole during the time $t$, and the linear space $s = pt$ (§§ 57 and 111), which is right through vertically. For example, during the first second the parallotipole has rotated through an angle $\alpha = \theta t_1 = \frac{3}{8} \times 1.566 \times 1 = 7.68$ units, $m$-measure, i.e., $(7.68 + 2\pi) = 1.92$ revolutions, while $P$, has sunk through $s = pt_1 = 3.84$ ft., vertically. The tension in the cord, from (2), is $$P = (48(1 - 1.566 \times \frac{3}{8} + g)) = (48(1 - 0.94)) = 36.48 \text{ lbs}.$$ The pressures at the bearings will be as follows, at any instant: from (4) and (8), $X$, and $X'$ must individually be zero; from (6) $Z_2 = Q' = PV_y = 90$ lbs.; while from (5) and (7), $Y_x = m - 2.28$ lbs., and $Y'_x = -15.30$ lbs., and should point in a direction opposite to that in which they were assumed in Fig. 129 (see last lines of § 39). **116. Torsion Balance. A Variable Speed Rotary Motion.** **Axis Fixed—A homogeneous solid having as axis of symmetry is suspended by an elastic prism, or filament (whose mass may be neglected), so that the latter is vertical and coincident with the axis of symmetry, and is not only supported, but prevented from turning at its upper extremity. If the solid is turned about its axis away from its position of rest and set free, the torsional** A diagram showing a torsion balance with a homogeneous solid suspended by an elastic prism or filament. Fig. 129 DYNAMICS OF A RIGID BODY. 117 elasticity of the rod or filament, which is fixed in the solid, causes an oscillatory rotary motion. Required the duration of an oscillation. Fig. 130. Take the axis $Y$ at the middle of the oscillation (the original position of rest). Backward the time from the instant of passing this position. Let the initial angular velocity be $\omega$. As the motion increases as $\omega$ diminishes, $\theta$ is negative. To consider the body free, conceive the rod cut close to the body (in which it is firmly inserted), and in the section thus exposed put in the vertical tension $P$, and also the horizontal forces forming a couple to which at any instant the twisting action (of the portion of rod removed upon the part left in the free body) is known to be due. Call the moment of this couple $Qh$ (known as the moment of torsion); it is variable, being directly proportional to the angle $\alpha$; hence, by experiment it is found that $Qh = C\theta$, when $a_1 = a_2$, for any value of $\theta$ it will be $Qh = (Qh + a_1)a = Ca$, in which $C$ is the constant factor. At any instant, therefore, the forces acting are $G$, $P$, and those equivalent to the couple whose moment $= Qh = Ca$. (No lateral support is required; the student would find the $X_r$, $X_v$, and $Y_v$ of Fig. 129 to be individually zero, if put in; remembering that here, $\dot{x}$ and $\dot{y}$ both = 0, as also $fdM_{xx}$ and $fdM_{yy}$; and that the forces of the couple will not be represented in any of the six summations of § 114, except in $\dot{x}$ and $\dot{y}$.) From eq. (XIV.), § 114, we have $-Qh$, i.e., $-Ca = \theta F_x$ from which $$\theta = -\left(C + f_x\right)\theta,$$ or, for short, $$\theta = -Ba.$$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) Since $B$ is constant, and there is an initial (angular) velocity $\omega_0$, and since the variables $\theta$, $\omega$, and $a$, in angular motion correspond precisely to those $(x,v,y,\dot{v})$ of rectangular motion, it is evident that the present case of harmonic motion, already discussed in Problem 2 of § 59. Applying the results there obtained, since $B$ of eq. (1) corresponds to the $a$ of that problem, we find that the oscillations are *isochronal*, i.e., their 118 MECHANICS OF ENGINEERING. duration are the same whatever the amplitude (provided the elasticity of the rod is not impaired), and that the duration of one oscillation (from one extreme position to the other) is $$t = \pi + \sqrt{B}$$ or finally, $$t = \pi + \sqrt{a_1 l_2 + Qb_3} \dots \dots \dots (2)$$ 117. The Compound Pendulum is any rigid body allowed to oscillate without friction under the action of gravity when mounted on a horizontal axis. Fig. 131 shows the body free, in any position during the progress of the oscillation. $C$ is the centre of gravity; let $\overline{OC} = s$. From (XIV), § 114, we have $x$ (mom. about fixed axis) $$\therefore - Gs \sin \alpha = OI_x$$ and $\theta = -Gs \sin \alpha + I_x = -Mgs \sin \alpha + MI_x$$ i.e., $\theta = -gs \sin \alpha + k^*_{\theta}$. (1) Hence $\theta$ is variable, proportional to $\sin \alpha$. Let us see what the length $l = OK$, of a simple circular pendulum, must be, to have at this instant (i.e., for this value of $\alpha$) the same angular acceleration as the rigid body. The linear (tangential) accelerations of $K$, the extremity of the required simple pendulum would be (§ 77) $p_t = -g \sin \alpha$, and hence its angular acceleration would $= -g \sin \alpha + l$. Writing this equal to $\theta$ in eq. (1), we obtain $$l = k^*_{\theta} + l$$ But this is independent of $\alpha$; therefore the length of the simple pendulum having an angular acceleration equal to that of the oscillating body is the same in all positions of the latter, and if the two begin to oscillate simultaneously from a position of rest at any given angle $\alpha$, with the vertical, they will keep abreast of each other during the whole motion, and hence have A diagram showing a compound pendulum with a body free to oscillate horizontally on a horizontal axis. Fig. 131 DYNAMICS OF A RIGID BODY. 118 the same duration of oscillation; which is . . . , for small amplitudes (§ 78), $$\ell = \pi \sqrt{1 + g} = \pi \sqrt{k_s^2 + g s} \quad \ldots \quad (3)$$ \(K\) is called the centre of oscillation corresponding to the given centre of suspension \(O\), and is identical with the centre of percussion (§ 115). Example. Required the time of oscillation of a cast-iron cylinder, whose diameter is 2 in. and length 10 in., if the axis of suspension is taken 4 in. above its centre. If we use 39.3 for \(g\), all linear dimensions should be in feet and times in seconds. From § 100, we have $$I_c = M(4^2 + 4^2) = M(4 \cdot 16 + 4 \cdot 16) = M_{16} \cdot 16$$ From eq. (3), § 88, $$I_c = I_s + Ms' = M(4 \cdot 4 + 4 \cdot 4) = M \cdot 0.170;$$ :. \(k_s^2 = 0.170\) sq. ft.; :. \(t = \pi \sqrt{0.170 + (32 \times 4)} = 0.395\) sec. 118. The Centres of Oscillation and Suspension are Interchangeable. (Strictly speaking, these centres are points in the line through the centre of gravity perpendicular to the axis of suspension.) Refer the centre of oscillation \(K\) to the centre of gravity, thus (Fig. 182, at (1)): $$s_t = t - s = \frac{M k_s^2}{Ms'} - s = \frac{M k_s^2}{Ms} - s = k_s^2 - s.$$ (1) Now invert the body and suspend it at \(K'\); required \(CK'\), or \(s_t\), to find the centre of oscillation corresponding to \(K'\) as centre of suspension. By analogy from (1) we have $$s_t = k_s'^2 + s_t'; but from (1), k_s'^2 + s_t' = s_t^2;$$ :. \(s_t' = x\); in other words, \(K'_s\) is identical with \(O\). Hence, by inversion, Advantage may be taken of this to determine the length \(L\) of the theoretical simple pendulum vibrating seconds, and thus finally the acceleration of gravity from formula (3), § 117, viz, Fig. 182 120 MECHANICS OF ENGINEERING. when $t' = 1.0$ and $l$ (now = $L$) has been determined experimentally, we have $$g \text{ (in ft. per sq. second)} = L \text{ (in ft.)} \times \pi^2 \cdot \cdot \cdot (2)$$ This most accurate method of determining $g$ at any locality requires the use of a bar of metal, furnished with a sliding weight for shifting the centre of gravity, and with two projecting blocks provided with knife-edges. These blocks can also be shifted and clamped. By suspending the bar by one knife-edge on a proper support, the duration of an oscillation is computed by counting the total number in as long a period of time as possible; it is then reversed and suspended on the other with like observations. By shifting the blocks between successive experiments, the duration of the oscillation in one position is made the same as in the other, i.e., the distance between them is such that the bar is in equilibrium vibrating in the computed time (if the knife-edges are not equidistant from the centre of gravity), and is carefully measured. The $l$ and $t'$ of eq. (3), § 117, being thus known, $g$ may be computed. Professor Bartlett gives as the length of the simple pendulum vibrating seconds at any latitude $\beta$ $$L \text{ (in feet)} = 3.8058 - 0.008318 \cos 2\beta.$$ 119. Isochronal Axes of Suspension.—In any compound pendulum, for any axis of suspension, there are always three others, parallel to it in the same gravity-plane, for which the oscillations are made in the same time as for the first. For any assigned time of oscillation $t'$, eq. (3), § 117, compute the corresponding distance $CO = x$ of $O$ from $C$; i.e., from $$t'^2 = \pi^2 \frac{M_1k_1}{M_{gs}} + \frac{M_2k_2}{M_{gs}},$$ we have $$x = (\pi^2 t'^2 - 2) \pm (\sqrt{\pi^2 t'^2 - 4}) - k_1 \cdot \cdot \cdot (1)$$ Hence for a given $t'$, there are two positions for the axis $O$ parallel to any axis through $C$, in any gravity-plane, on both sides; i.e., four parallel axes of suspension, in any gravity-plane, giving equal times of vibration; for two of these axes DYNAMICS OF A RIGID BODY. 121 we must reverse the body. E.g., if a slender, homogeneous, prismatic rod be marked off into thirds, the (small) vibrations will be of the same duration, if the centre of suspension is taken at either extremity, or at either point of division. Example.—Required the positions of the axes of suspension, parallel to the base, of a right cone of brass, whose altitude is six inches, radius of base one inch, and weight which incline is 0.304 lbs., so that the time of oscillation may be half-second. (N.B. For variety, use the inch-pound-second system of units, first consulting § 51.) 120. The Fly-Wheel in Fig. 133 at any instant experiences a pressure $P$ against its crank-pin from the connecting-rod and a resisting pressure $P'$ from the teeth of a spur-wheel with A diagram showing a flywheel with teeth and a connecting rod. The diagram includes labels such as P, P', m, n, r, and various other symbols. Fig. 133. which it gears. Its weight $G$ acts through $C$ (nearly), and there are pressures at the bearings, but these latter and $G$ have no moments about the axis $C$ (perpendicular to paper). The figure shows it free. $P'$ being assumed constant (in practice this depends on the resistance met by the machines which $D$ drives, and the fluctuation of velocity of their moving parts). $P'$, and therefore $T$ its tangential component, are variable, depending on the effective steam-pressure on the piston at any instant, on the obliquity of the connecting-rod, and in high-speed engines on the masses and motions of the piston and connecting-rod. Let $r$ = radius of crank-pin circle, and $\alpha$ = perpendicular from $C$ on $P'$. From eq. (XIV.), § 114, we have $$Tr - P'a = \theta T_{a} \cdot \theta = (Tr - P'a) + I_{a} \cdot \quad (1)$$ 192 MECHANICS OF ENGINEERING. as the angular acceleration at any instant; substituting which in the general equation (VIII), § 10, we obtain $$I_{\omega_{\infty}} = Trda - P^{\prime}ada \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ From (1) it is evident that if at any position of the crank-pin the variable $Tr$ is equal to the constant $P^{\prime}a$, $\theta$ is zero, and consequently the angular velocity $\omega$ is either a maximum or a minimum. Suppose this is known to be the case both at $m$ and $n$; i.e., suppose $T$, which was zero at the dead-point $A$, has been gradually increasing, till at $n$, $Tr = P^{\prime}a$; and thereafter increased until at $m$, $Tr$ again = $P^{\prime}a$, and continues to diminish toward the dead-point $B$. The angular velocity $\omega$, whatever it may have been on passing the dead-point $A$, diminishes, since $\theta$ is negative, from $A$ to $n$, where it is $\omega_{\infty}$ a minimum; increases from $n$ to $m$, where it reaches a maximum value, $\omega_{\infty}$, and $n$ and $m$ being known points, and supposing $\omega_{\infty}$ known, let us inquire what $\omega_{\infty}$ will be. From eq. (2) we have $$I_{\omega}\int_{m}^{n} ad\omega = \int_{n}^{m} Trda - P^{\prime}\int_{n}^{m} ad\alpha \quad . . . . . . . . . . . . . . . . . . . (3)$$ But $d\alpha = d\theta$ is an element of the path of the crank-pin, and also the "virtual velocity" of the force $T$, and $ad\alpha = d\theta'$, an element of the path of a point in the pitch-circle of the fly-wheel, the small space through which $P^{\prime}$ is overcome in $dt$. Hence (3) becomes $$I_{\omega}\left(\omega_{\infty} - \omega_{\infty}\right) = \int_{A}^{B} Td\theta - P^{\prime}\times \text{linear arc } EF. \quad (4)$$ To determine $\int_{A}^{B} Td\theta$ we might, by a knowledge of the varying steam-pressure, the varying obliquity of the connecting-rod, etc., determine $T$ for a number of points equally spaced along the curve $mn$, and obtain an approximate value of this sum by Simpson's Rule; but a simpler method is possible by noting (see eq. § 65) that each term $Td\theta$ of this sum = the corresponding term $Pdz$ in the series $\sum dz$, in which $P =$ the DYNAMICS OF A RIGID BODY. 123 effective steam-pressure on the piston in the cylinder at any instant, $dz$ the small distance described by the piston while the crank-pin describes any $da$, and $n'$ and $m'$ the positions of the piston (or of cross-head, as in Fig. 133) when the crank-pin is at $a$ and $m$ respectively. (4) may now be written $$I_{\text{ch}}(\omega_0 - \omega_*) = \int_{a}^{m'} P dz - F'' + \text{linear arc} = EP, \quad (5)$$ from which $\omega_*$ may be found as proposed. More generally, it is available, alone (or with other equations), to determine any one (or more, according to the number of equations) unknown quantity. This problem, in rotary motion, is analogous to that in § 59 (Prob. 4) for rectilinear motion. Friction and the inertia of piston and connecting-rod have been neglected. As to the time of describing $da$ are now from equations similar to (5), we shall employ Simpson's Rule for this purpose, distilling it into an even number of equal parts, calling them $dt$, etc., and then employ Simpson's Rule for an approximate value of the sum $$\sum_{n=1}^N f^n da$$ (from eq. (VI.), § 110); e.g., with four parts, we would have $$t = \frac{1}{16} (\angle nCm, \pi \text{mean}) \left[ \frac{1}{\omega_0} + \frac{4}{\omega_1} + \frac{4}{\omega_2} + \frac{1}{\omega_N} \right] dt, \quad (6)$$ 191. Numerical Example. Fly-Wheel.—(See Fig. 133 and the equations of § 120.) Suppose the engine is non-condensing and non-expansive (i.e., that $P$ is constant), and that $$P = 5500 \text{ lbs}, \quad a = 6 \text{ in.}, \quad b = 4 \text{ ft}, \quad a = 2 \text{ ft},$$ and also that the wheel is to make 192 revolutions per minute, i.e., that its mean angular velocity is to be $$\omega' = 48^\circ \times 2\pi, \text{i.e., } \omega' = 48^\circ \text{ radians per sec}. $$ First, required the amount of the resistance $P''$ (constant) that there shall be no permanent change of speed, i.e., that the angular velocity shall have the same value at the end of a complete revolution as at the beginning. Since an equation of the form of eq. (6) holds good for any range of the motion, let 124 MECHANICS OF ENGINEERING. that range be a complete revolution, and we shall have zero as the left-hand member; $fPdz = P \times 2\ ft. = 5500$ lbs. $\times 2\ ft.$, or 11,000 foot-pounds (as it may be called); while $P^{\prime}$ is unknown, and instead of lin, are $Ez$ we have a whole circumference of $2\ ft.$ radius, i.e., $4\ ft.$ $\therefore 0 = 11,000 \times 4 \times 3.816$, whence $P^{\prime} = 575$ lbs. Secondly, required the proper mass to be given to the fly-wheel of $2\ ft.$ radius that in the forward stroke (i.e., while the crank-pin is describing its upper semicircle) the max. angular velocity $\omega_{a}$ shall exceed the minimum $\omega_{b}$ by only $\frac{1}{n}\omega'$, assuming (which is nearly true) that $\frac{1}{n}(\omega_{a} + \omega_{b}) = \omega'$. There being now three unknowns, we require three equations, which are, including eq. (5) of § 120, viz: $$M_{e}^{(n)}(\omega_{a} + \omega_{b})(\omega_{a} - \omega_{b}) = \int_{n}^{m}Pdz - P^{\prime} \times linear \ are EF; (6)$$ $$\frac{1}{n}(\omega_{a} + \omega_{b}) = \omega' = 4\pi; (7) \quad and \quad \omega_{a} - \omega_{b} = \frac{1}{n}\omega' = 4\pi. (8)$$ The points $n$ and $m$ are found most easily and with sufficient accuracy by a graphic process. Laying off the dimensions to scale, by trial such positions of the crank-pin are found that $T$, the tangential component of the thrust $P$ produced in the connecting-rod by the steam-pressure $P$ (which may be resolved into two components acting along the axis of the rod and at right angles to itself) is $(a + r)\sqrt{P'}$, where $r = 3500$ lbs. These points will be $n$ and $m$ (and two others on the lower semicircle). The positions of the piston $n'$ and $m'$ corresponding to $n$ and $m$ of the crank-pin, are also found graphically in an obvious manner. We thus determine the angle $COn$ to be $100^\circ$, so that linear arc $EF = 4\pi\ ft. \times 2\ ft. = 8\pi\ ft.$, while $$\int_{n'}^{m'}Pdz = 5500 \text{ lbs.} \times \int_{n'}^{m'}ds = 5500 \times n'm' = 5500 \times 0.77\ ft.,$$ $n'm'$ being scaled from the draft. Now substitute from (7) and (8) in (5), and we have, with $k_2 = 2\ ft.$ (which assumes that the mass of the fly-wheel is concentrated in the rim), $$M_{e}^{(n)}(\omega_{a} + \omega_{b})(\omega_{a} - \omega_{b}) = E(4\pi)^{2}.$$ DYNAMICS OF A RIGID BODY. 125 $(G+g)\times4\times4\pi\times4\pi=5500\times0.77-875\times4\pi,$ which being solved for $G$ (with $g=38.3$; since we have used the foot and second), gives $G=600.7$ lb. The points of max. and min. angular velocity on the back-stroke may be found similarly, and their values for the fly-wheel are $\omega_{b}$ and $\omega_{a}$, respectively, slightly from the $\omega_{b}$ and $\omega_{a}$ of the forward stroke. Professor Cotterill says that the rim of a fly-wheel should never have a max. velocity > 80 ft. per sec.; and that if made in segments, not more than 40 to 50 feet per second. In the present example we have for the forward stroke, from eqn. (7) and $(b)$, $\omega_{b}=13.5$ (w-measure units) per second; i.e., the corresponding velocity of the wheel-rim is $v_{a}=\omega_{b}d=26.4$ feet per second. 129. Angular Velocity Constant. Fixed Axis.--If $\omega$ is constant, the angular acceleration, $\theta$, must be zero at all times, which requires $\Sigma$ (mom.) about the axis of rotation to be $=0$ (eq. [XIV], § 114). An instance of this occurs when the only forces acting are those due to friction between the axis and the body's weight, parallel to or intersecting the axis; the values of these reactions are now to be determined for different forms of bodies, in various positions relatively to the axis. (The opposites and equals of these reactions, i.e., the forces with which the axis acts upon the bearings, are sometimes stated to be due to the "centrifugal force," or "centrifugal action," of the revolving body.) Take the axis of rotation for Z, then, with $\theta=0$, the equations of § 114 reduce to $$\begin{array}{lll} \Sigma X &=& -\omega^2Mz & \dots\dots & (\mathrm{IXa}) \\ \Sigma Y &=& -\omega^2My & \dots\dots & (\mathrm{Xa}) \\ \Sigma Z &=& 0 & \dots\dots & (\mathrm{Xla}) \\ \Sigma \text{ mom}_x &=& -\omega^2/Myz & \dots\dots & (\mathrm{XIIa}) \\ \Sigma \text{ mom}_y &=& +\omega^2/Mxz & \dots\dots & (\mathrm{XIIIa}) \\ \Sigma \text{ mom}_z &=& 0 & \dots\dots & (\mathrm{XIVA}) \\ \end{array}$$ 136 MECHANICS OF ENGINEERING. For greater convenience, let us suppose the axes $X$ and $Y$ (since their position is arbitrary so long as they are perpendicular to each other and to $Z$) to revolve with the body in its uniform rotation. 129a. If a homogeneous body have a plane of symmetry and rotate uniformly about any axis $Z$ perpendicular to that plane (intersecting it at $O$), then the acting forces are equivalent to a single force, = $\omega^2 M_0$, applied at $O$ and acting in a gravity-line, but directed away from the centre of gravity. It is evident that such a force $P = \omega^2 M_0$, applied as stated (see Fig. 138), will satisfy all the conditions expressed in the preceding equation, viz., $X$ through the centre of gravity, so that $x = \frac{d}{dx}$. For each ($LX.A$) $P$ must = $\omega^2 M_0$, while in each of the other summations the left-hand member will be zero, since $P$ lies in the axis of $X$; and as their right-hand members will also be zero for the present body ($y = 0$; and each of the sums $fd.M_y$ and $fd.M_z$ is zero, since for each term $df.M_y + x$ there is another $df.M_y - x$ to cancel it; and similarly, for $fd.M_z$, they are also satisfied; Q.E.D.* Hence a single point of support at $O$ will suffice to maintain the uniform motion of the body, and the pressure against it will be equal and opposite to $P$. First Example.—Fig. 138. Supposing (for greater safety) that the fly-wheel revolves at 150 revolutions per minute of each segment of a fly-wheel is maintained solely by the tension in the corresponding arm, $P$, required the value of $P$ if the segment and arm together weigh $\frac{d}{dx}$ of a ton, and the distance of their centre of gravity from the axis is $\rho = 20$ in., i.e., = $\frac{4}{3}$ ft. With the foot-ton-second system of units, with $\rho = 32.3$, we have $$P = \omega^2 M_0 = [2\pi \times 2\pi] \times [2\pi + 32.3] \times \frac{4}{3} = 0.83 \text{ tons},$$ or 160 lbs. * That is, neglecting gravity. Fig. 138. DYNAMICS OF A RIGID BODY. 127 **Second Example.**—Fig. 137. Suppose the uniform rotation of the same fly-wheel depends solely on the tension in the rim, required its amount. The figure shows the half-rim free, with the two equal tensions, $P$, put in at the surfaces exposed. Here it is assumed that the arms exert no tension on the rim. From § 122a we have $2P = \omega^2 M$, where $M$ is the mass of the half-rim, and $\rho$ its gravity co-ordinate which may be obtained approximately by § 26, Problem 1, considering the rim as a circular wire, viz., $\rho = 2r + \pi$. Let $M = (180 \text{ lbs.)} / r$, with $r = 2$ ft. We have then $$P = \frac{1}{2}(9)(180 + 32.5\pi) = 1718 \text{ lbs.}$$ (In reality neither the arms nor the rim sustain the tensions just computed; in treating the arms we have supposed no duty done by the rim, and vice versa. The actual stresses are less, and depend on the yielding of the parts. Then, too, we have supposed the wheel to take no part in the transmission of motion by belting or gearing, which would cause a bending of the arms, as shown in Fig. 138.) **123b.** If a homogeneous body have a line of symmetry and rotate uniformly about an axis parallel to it (O being the foot of the perpendicular from the centre of gravity on the axis), then the acting forces are equivalent to a single force $P = \omega^2 M$, applied at $O$ and acting in a gravity-line away from the centre of gravity. Taking the axis $X$ through the centre of gravity, $Z$ being the perpendicular to $XY$, Fig. 138, which $Z$ is the line of symmetry, pass an auxiliary plane $ZY$ parallel to $ZY$. Then the sum $f dMz$ may be written $fd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ which is $gd(Mz + x'dx)$ = 0, and every term \(dM(x' - z)\) cancelled by a numerically 128 MECHANICS OF ENGINEERING equal term $dM(-x')$ of opposite sign. Hence $fdM_{xx}=0$. Also $fdM_{yy}=0$, since each positive product is annulled by an equal negative one (from symmetry about $YZ$). Since, also, $y=0$, all six conditions in § 122 are satisfied. Q. E. D. If the homogeneous body is any solid of revolution whose geometrical axis is parallel to the axis of rotation, the foregoing is directly applicable. 123a. If a homogeneous body revolve uniformly about any axis lying in a plane of symmetry, the acting forces are equivalent to a single force $F'=\omega dM$, acting parallel to the gravitational which is perpendicular to the plane ($XZ$), and away from the centre of gravity, its distance from any origin $O$ in the axis $Z$ being = $\sqrt{(\omega^2r^2+g^2)}+\sqrt{g^2}$ (the plane $XZ$ being a gravity plane).—Fig. 130. From the position of the body we have $F'=\omega dM$, and $y=0$; hence if a value $\omega dM$ be given to $P$ and it be made to act through $Z$ and parallel to $X$, and away from the centre of gravity, all the conditions of § 122 are satisfied except (XIIa.) and (XIIla.). But symmetry about the plane $XZ$ makes $fdM_{zz}=0$, and satisfies (XIIa.), and by placing $F'$ at a distance $a=f'dMz+M_0$ from $O$ along $Z$ we satisfy (XIIla.). Q. E. D. Example.—A slender homogeneous, prismatic rod, of length $l$, is to have a uniform motion, about a vertical axis passing through one extremity, maintained by a cord-connection with a fixed point in this axis. Fig. 140. Given $\omega$, $\varphi$, $t$, $(\bar{\rho}=\frac{1}{2}\cos\varphi)$, and $F'$ the cross-section of the rod, let $s=$ the distance from $O$ to any $dM$ of the rod. $dM=F'yds+\frac{1}{2}g^2s$. The sum of any $dM=s\cos\varphi$; its sum $\sin\varphi$:, $\therefore fdM_{zz}=(F'y+\frac{1}{2}g^2)\sin\varphi\cos\varphi\int s'ds$ $=4(F'y+\frac{1}{2}g^2)\sin\varphi\cos\varphi=\frac{1}{2}Mr^2\sin\varphi\cos\varphi.$ Fig. 130. Fig. 140. DYNAMICS OF A RIGID BODY. 129 Hence $a = \frac{fdMz}{\omega M} = \frac{d}{dt}\sin\varphi$, and the line of action of $P' = \omega'Mz = \omega'(fyl + g)\cos\varphi$ is therefore higher up than the middle of the rod. Find the intersection of $G$ and the horizontal drawn through $Z$ at distance $a$ from $O$. Determine the magnitude of the parallel force $GP'$ attaching the rod so as to make it coincide with $P'$, for this will satisfy the condition of maintaining the motion, when once begun, viz., that the acting forces $G$, and the cord-tension $P'$, shall be equivalent to a force $P = \omega'Mp$, applied horizontally through $Z$ at a distance $a$ from $O$. 123. Free Axes. Uniform Rotation.—Referring again to § 122 and Fig. 134, let us inquire under what circumstances the lateral forces, $X_x$, $Y_x$, $X_y$, $Y_y$, with which the bearings press the axis, to maintain the motion, are individually zero, i.e., that the bearings are not needed, and may therefore be removed (except a smooth horizontal plane to sustain the body's weight), leaving the motion undisturbed like that of a top " asleep". For this, not only must $\frac{d}{dt}(x + y)$ and $\frac{d}{dt}(y - x)$ both equal zero (since either $\frac{d}{dt}(x + y) = 0$, or $\frac{d}{dt}(y - x) = 0$), but also $X_x$ and $Y_x$ (similarly) $X_y$ (moms.)$_x$ and $Z$ (moms.)$_y$ must each be zero. The necessary peculiar distribution of the body's mass about the axis of rotation, then, must be as follows (see the equations of § 129): First, $\bar{x}$ and $\bar{y}$ each be 0, i.e., the axis must be a gravity-axis. Secondly, $\frac{d}{dt}Mx = 0$, and $\frac{d}{dt}My = 0$, the origin being anywhere on the axis of rotation. An axis ($Z$) of a body failing these conditions is called a Free Axis. It is shown that one of the three Principal Axes for the centre of gravity (see § 107) be made an axis of rotation (the other two being taken for $X$ and $Y$), the conditions $\bar{x} = 0$, $\bar{y} = 0$, $\frac{d}{dt}Mx = 0$, and $\frac{d}{dt}My = 0$, are all satisfied; it follows that every rigid body has at least three free axes, which are the Principal Axes of Inertia of the centre of gravity at right angles to each other. In the case of homogeneous bodies free axes can often be determined by inspection : e.g., any diameter of a sphere; any 130 MECHANICS OF ENGINEERING. transverse diameter of a right circular cylinder through its centre of gravity, as well as its geometrical axis; the geometrical axis of any solid of revolution; etc. 184. Rotation about an Axis which has a Motion of Translation. ---Take only the particular case where the moving axis is a gravity-axis. At any instant, let the velocity and acceleration of the axis be $v$ and $p$; the angular velocity and acceleration about that axis, $\omega$ and $\theta$. Then, since $dM/dt = dM\omega dp + p = d^2Mdt$, we may, by forming imaginary equivalents system in Fig. 184, take each $dM$ to move with the simultaneous action of $dP = dMdp$ parallel to $X$, of the tangential $dT = dM\omega p$, and of the normal $dN = dM(\omega dp + p) = d^2Mdt$. Take $X$ in the direction of translation, $Z$ (perpendicular to paper through $O$) is the moving gravity-axis; $Y$ perpendicular to both. At any instant we shall have, then, the following conditions for the acting forces (remembering that $\sin \phi = y$, $f_dM_y = M_y = 0$; etc.): $$X = f_dP - f_dT \cos \phi - f_dN \cos \phi = M_p; \quad (1)$$ $$Y = f_dT \cos \phi - f_dN \sin \phi = 0; \quad (2)$$ $$Z_momk = f_dTp - f_dPy = 0; dMp = 0; dNp = 0.$$ (3) and three other equations not needed in the following example. **Example—A homogeneous solid of revolution rolls (without slipping) down a rough inclined plane. Investigate the motion.** Considering the body free, the acting forces are $G$ (known) and $N$ and $P$, the unknown normal and tangential components of the action of the plane on the roller. If slipping occurs, then $P$ is the sliding friction due to the pressure $F_1$ ($\S 156$); here, however, it is A diagram showing a rolling object on an inclined plane with forces acting on it. 185 A diagram showing a rolling object on an inclined plane with forces acting on it. 186 DYNAMICS OF A RIGID BODY. less by hypothesis (perfect rolling). At any instant the four unknowns are found by the equations $$\Sigma X,\mathrm{i.e.,}\ G\sin\beta-P=(G+g)p;\quad(1)$$ $$\Sigma Y,\mathrm{i.e.,}\ G\cos\beta-N=0;\quad(2)$$ $$\Sigma\mathrm{mom}_{z},\mathrm{i.e.,}\ Pa=\theta Mz^{2};\quad(3)$$ while on account of the perfect rolling, $$\theta_{a}=p.\quad(4)$$ Solving, we have, for the acceleration of translation, $$p=g\sin\beta+[1+(k_{2}+a^{\prime})].$$ (If the body aid without friction, $p$ would $=g\sin\beta$.) Hence for a cylinder ($§ 97$), $k_{2}$ being $=k$, we have $p=g\sin\beta$; and for a sphere ($§ 108$) $p=g\sin\beta$. (If the plane is so steep or so smooth that both rolling and slipping occur, $\theta_{a}$ no longer $=p$, but the ratio of $P$ to $N$ is known from experiments on sliding friction; hence there are still four equations.) The motion of translation being thus found to be uniformly accelerated, we may use the equations of § 56 for finding distance, time, etc. **Query.**—How may we distinguish two spheres by allowing them to roll down the same inclined plane, if one of them is silver and solid, while the other is of gold, but silvered and hollow, so as to be the same as the first in diameter, weight, and appearance? 125. Parallel-Rod of a Locomotive.—When the locomotive moves uniformly, each $dM$ of the rod between the two (or three) driving wheels has a uniform velocity about a centre of its own on the line $BD$, Fig. 148, and with a velocity $v$ and radius $r$ common to all, and likewise has a horizontal uniform motion of translation. Hence if we inquire what are the reactions $P$ of its supports, as induced solely by its weight and motion, when in its lowest position (independently of any thrust along 131 A diagram showing a parallel-rod of a locomotive with a uniform motion of translation. 148 132 MECHANICS OF ENGINEERING. the rod), we put $\mathbb{X}Y$ of (I.) = $\mathbb{X}Y$ of (II.) (II. shows the imaginary equivalent system), and obtain $$2P - G = f dN = f dM e^a + r = (e^a + r) / dM = M e^a + r.$$ **Example.** Let the velocity of translation be 50 miles per hour, the radius of the pins be 18 in. = 4 ft., and $r$ = half that of the driving-wheel, while the weight of the rod is 200 lbs. With $g = 32.2$, we must use the foot and second, and obtain $$v = \frac{1}{4}(50 \times 5980 + 3600) \text{ ft. per second} = 36.6;$$ while $$M = 200 + 32.2 = 300 \times 0.0310 = 6.30;$$ and finally $$P = \frac{1}{4}[300 + 6.3(36.6)^2 + g] = 2868.3 \text{ lbs.,}$$ or nearly 14 tons, about thirty times that due to the weight alone. **130.** So far in this chapter the motion has been prescribed, and the necessary conditions determined, to be fulfilled by the acting forces at any instant. Problems of a converse nature, i.e., where the initial state of the body and the acting forces are given while the resulting motion is required, are of much greater interest than those considered above. The reader is referred to Rankine's Applied Mechanics. A treatment of the Gyroscope will be found in the American Journal of Science for 1887, and in the article of that name in Johnson's Cyclopaedia. WORK, ENERGY, AND POWER. 133 CHAPTER VI. WORK, ENERGY, AND POWER 197. Remark.—These quantities as defined and developed in this chapter, though compounded of the fundamental ideas of matter, force, space, and time, enter into theorems of such wide application and practical use as to more than justify their consideration as separate kinds of quantity. 198. Work in a Uniform Translation. Definition of Work— Let Fig. 144 represent a rigid body having a motion of translation parallel to $X$, acted on by a system of forces $P$, $P_n$, $R_n$ and $R$, $\vec{r}_n$ which remain constant. Let $s$ be any distance described by the body during its motion; then $2X \cdot s$ must be zero (§ 109), i.e., noting that $R_x$ and $R_y$ have negative $X$ components (the supplements of their angles with $X$ are used), $$P_x \cos \alpha + P_n \cos \alpha - R_x \cos \alpha - R_y \cos \alpha = 0;$$ or, multiplying by $s$ and transposing, we have (noting that $s \cos \alpha_s = x$, the projection of $s$ on $P_x$, that $s \cos \alpha_s = y$, the projection of $s$ on $P_y$, and so on). $$P_{\phi} + P_{\phi_n} = R_{\phi} + R_{\phi_n}. \quad . . . . . . . . . . . . . . . . . . . . . . . . . . (a)$$ The projections $s_x$, $s_y$, etc., may be called the distances described in their respective directions by the forces $P_x$, $P_n$, etc.; $P$, and $P$, having moved forward, since $x_s$ and $y_s$ fall in front of the initial position of their point of application; $R_x$ and $R_y$, backward, since they fall behind the point of application in their case. (By forward and backward we refer to the direc... Fig. 144 134 MECHANICS OF ENGINEERING. tion of each force in turn.) The name work is given to the product of a force by the distance described in the direction of the force by the point of application. If the force moves forward (see above), it is called a working force, and is said to do the work (e.g., $P_{A}$) expressed by this product; while if backward, it is called a resistance, and is then said to have the work (e.g., $R_{A}$) done upon it, in overcoming it through the distance mentioned (it might also be said to have done negative work). Eq. (a) above, then, proves the theorem that : In a uniform translation, the working forces do an amount of work which is entirely applied to overcoming the resistance. 129. Unit of Work.—Since the work of a force is a product of force by distance, it may logically be expressed as so many foot-pounds, inch-pounds, kilogram-meters, according to the system of units employed. The ordinary English unit is the foot-pound, or ft.-lb. It is of the same quality as a force-moment. 130. Power.—Work as already defined does not depend on the time occupied, i.e., the work $P_{A}t$, is the same whether performed in a long or short time; but the element of time is of so great importance in all the applications of dynamics, as well as in each practical commercial matter as water-supply, consumption of fuel, fatigue of animals, etc., that the rate of work is a consideration both of interest and necessity. Power is the rate at which work is done, and one of its units is frequently used in English practice; a larger one will be mentioned presently. The power exerted by a working force, or expended upon a resistance, may be expressed symbolically as $$L = P_A t \text{ or } R_A t,$$ in which $t$ is the time occupied in doing the work $P_A t$, or $R_A t$, (see Fig. 144); or if $\nu$, is the component in the direction of the force $P$, of the velocity $\nu$ of the body, we may also write $$L = P_A \nu. \quad \quad \quad \quad \quad \quad \quad (b)$$ **WORCE, ENERGY, AND POWER.** 135 **131. Example.—Fig. 145, shows as a free body a sledge which is being drawn uniformly up a rough inclined plane by a cord parallel to the plane. Required the total power exerted (and expended), if the tension in the cord is $P = 100$ lbs., the weight of sledge $R_1 = 160$ lbs., $\beta = 30^\circ$, and the sledge rises 240 ft. each minute. $N$ and $R_2$ are the normal and parallel (i.e., $R_2$ = friction) components of the reaction of the plane on the sledge. From eq. (1), § 128, the work done while the sledge advances through $s = 240$ ft. may be obtained either from the working forces, which in this case are represented by $P$, alone, or from the resistances $R_1$ and $R_2$. Take the former method first. Projecting $s$ upon $P$, we have $s_p = s$. Hence $P \cdot s_p = 100 \times 240 = 24,000$ ft-lbs. of work done in 60 seconds. That is, the power exerted by the working force is $$L = P \cdot s_p + t = 400 \text{ ft-lbs. per second}.$$ As to the other method, we notice that $R_1$ and $R_2$ are resistances, since the projections $s_p = s \sin \beta$, and $s_p = s$, would fall back of their points of application in the initial position, while $N$ is neutral, i.e., neither a working force nor a resistance, since the projection of $s$ upon it is zero. From $$\Sigma X = 0 \Rightarrow - R_1 \sin \beta + P_1 = 0,$$ and from $$\Sigma Y = 0 \Rightarrow N - R_2 \cos \beta = 0;$$ whence $R_1$ is 20 lbs., and $N$ is 138.5 lbs. Also, since $s_p = s \sin \beta = 240 \times \frac{1}{2} = 120$ ft., and $s_p = s$, we have for the work done upon the resistances (i.e., in overcoming them) in 60 seconds $$R_1 p_s + R_2 p_s = 160 \times 120 + 240 \times 240 = 24,000 \text{ ft-lbs},$$ and the power expended in overcoming resistance, $$L = 24,000 + 60 = 400 \text{ ft-lbs. per second},$$ as already derived. Or, in words the power exerted by the 136 MECHANICS OF ENGINEERING. Tension in the cord is expended entirely in raising the weight a vertical height of 2 feet, and overcoming the friction through a distance of 4 feet along the plane, every second; the motion being a uniform translation. 132. Horse-Power.--As an average, a horse can exert a tractive effort or pull of 100 lbs., at a uniform pace of 4 ft. per sec-ond, for ten hours a day without too great fatigue. This gives a power of 400 ft.-lbs. per second; but Boulton & Watt in rating their engines, and experimenting with the strong dray-horses of London, fixed upon 550 ft.-lbs. per second, or 83,000 ft.-lbs. per minute, as a convenient large unit of power. (The French horse-power, or cheval-pouvoir, is less than the English, being 75 kilogrammes per second, or 32,525 ft.-lbs. per minute.) The use of this unit has become common use. In the example in § 131, then, the power of 400 ft.-lbs. per second exerted in raising the weight and overcoming friction may be expressed as $(400+550)=\frac{1}{4}$ of a horse-power. A man can work at a rate equal to about $\frac{1}{4}$ of a horse-power, with proper intervals for eating and sleeping. 133. Kinetic Energy. Retarded Translation.--In a retarded translation of a rigid body whose mass $M$, suppose there are no working-forces, and that the resistances are constant and their resultant is $R$. (E.g., Fig. 146 shows such a case; a sledge, having an initial velocity $v$ and sliding on a rough horizontal plane, is gradually retarded by the friction $R$.) $R$ is paral- lel to the direction of translation (§ 109), and the acceleration is $a = - R/M$, hence from $v^2 = v_0^2 + 2ad$ we have $$v^2 = -(1 + M/R)v_0^2 \qquad \ldots \ldots \ldots (1)$$ But the projection of each do of the motion upon $R$ is $= -d$ itself; i.e. (§ 129). $Rdv$ is the work done upon $R$, in overcom- ing it through the small distance $ds$, and $fRdv$ is the sum of all such amounts of work throughout any definite portion of the motion. Let the range of motion be between the points Fig. 146 136 WORK, ENERGY, AND POWER. 137 where the velocity $v = c$, and where it = zero (i.e., the mass has come to rest). With these limits in eq. (1) (0 and $e'$ being the corresponding limits for $e$), we have $$\frac{Mv^2}{2} = \int e' Rde \qquad \qquad (c)$$ That is, in giving up all its velocity $v$ the body has been able to do the work $Rde$ (this, if $R$ remains constant, reduces to $Re'$) or its equal $\frac{Mv^2}{2}$. If, then, by energy we designate the ability to perform work, we give the name kinetic energy of a moving body to the product of its mass by half the square of its velocity ($\frac{Mv^2}{2}$); i.e., energy due to motion. (The antiquated term vis viva was once applied to the form $Mv^2$.) 134. Work and Kinetic Energy in any Translation. Let $P$ be the resultant of the working forces at any instant, $R$ that of the resistances; they (§109) will both act in a gravity-line parallel to the direction of translation. The acceleration at any instant is $p = (2X + M)\frac{dv}{dt}$, see 165. $$= (P - R) + M; \text{ hence from } vde = pde \text{ we have}$$ $$Mvde = Pde - Rdse \qquad \qquad (1)$$ Integrating between any two points of the motion as $O$ and $O'$ where the velocities are $v_0$ and $v'$, we have after transposition $$\int_{O'}^{O'} Pde = \int_{O'}^{O'} Rde + \left[\frac{Mv'^2}{2} - \frac{Mv_0^2}{2}\right] \qquad \qquad (d)$$ But $P$ being the resultant of $P_x$, $P_y$, etc., and $R$ that of $R_x$, $R_y$, etc., we may prove, as in §62, that if $du_x$, $du_y$, etc., be the respective projections of any de upon $P_x$, $P_y$, etc., while $dv_x$, $dv_y$, etc., are those upon $R_x$, $R_y$, etc., then $$Pde = P_x du_x + P_y du_y + \ldots \text{ and } Rdse = R_x dv_x + R_y dv_y, \ldots;$$ and $(d)$ may be rewritten 138 MECHANICS OF ENGINEERING. $$\int P.du + \int R.du + \ldots = \int R.du + \int R.du + \ldots + \left[ \frac{Mv^2}{2} - \frac{Mv^*}{2} \right]; (e)$$ or, in words: In any translation, a portion of the work done by the working force is applied in overcoming the resistances while the remainder equals the change in the kinetic energy of the body. It will be noted that the bracket in (e) depends only on the initial and final velocities, and not upon any intermediate values; hence, if the initial state is one of rest, and also the final, the total change in kinetic energy is equal to the work of the working force, which is entirely expended in the work of overcoming the resistances; but at intermediate stages the former exceeds the work so far needed to overcome resistances, and this excess is said to be stored in the moving mass; and as the velocity gradually becomes zero, this stored energy becomes available for aiding the working forces (which of themselves are then insufficient) in overcoming the resistances, and is then said to be restored. (The function of a fly-wheel might be stated in similar terms, but as that involves rotary motion it will be omitted.) Work applied in increasing the kinetic energy of a body is sometimes called "work of inertia," as also the work done by a moving body in overcoming resistances, and thereby losing speed. **135. Example of Steam-Hammer—Let us apply eq. (e) to determine the velocity $v'$ attained by a steam-hammer at the lower end of its stroke (the initial velocity being $=0$), just before delivering its blow upon a support, supposing that the stroke is 10 feet long, at each downward stroke is given by an indicator. Fig. 148. Weight of hammer mass is 392 lbs.; $M = 10$ (foot-pound-second system), $l = 1$ foot. The working force at any instant is $P' = G = 392$ lbs.; $P$, which is variable, but whose values at the several equally spaced A diagram showing a steam-hammer with a weight attached to a lever system. The lever has a pointer indicating the position of the weight. ```html
Diagram of a steam-hammer with a weight attached to a lever system.
``` WORK, ENERGY, AND POWER. 139 points $a$, $b$, $c$, $d$, $e$, $f$, $g$ are 800, 900, 900, 800, 800, 500, 450 lbs., respectively. $R$, the exhaust-pressure (16 lbs. per sq. inch $\times 20$ sq. inches piston-area) = 320 lbs., is the only resistance, and is constant. Hence from eq. (c), since here the projections $du_a$, etc., of any $da$ upon the respective forces are equal to each other and = $da$, $$P \cdot f_1 da + f_2 da = R_f \cdot da = \frac{Me^*}{2}.$$ (1) The term $f_2 da$ can be obtained approximately by Simpson's Rule, using the above values for six equal divisions, which gives $$\frac{1}{6}(800+4(900+800+500)) + 2(900+600)+4(500)$$ = 725 ft.-lbs. of work. Hence, making all the substitutions, we have, since $\int f_1 da = 1$ ft. $$320 \times 1 + 725 = 320 \times 1 + \frac{1}{2} Me^*; \therefore Me^* = 725 \text{ ft.-lbs.}$$ of energy to be expended in the forging. (Energy is evidently expressed in the same kind of unit as work.) We may then say that the forging receives a blow of 725 ft.-lbs. energy. The pressure actually felt at the surface of the hammer varies from instant to instant during the compression of the forging and the gradual stopping of the hammer, and depends on the readiness with which the metal yields. In this case we encountered is $R_{\text{max}}$ equal to depth of compression $s'$; we would have neglected the force of gravity, and noting that now the initial velocity is $v'$, and the final zero), from eq. (c), $$\frac{1}{2} M v'^2 = R_{\text{max}}'; \text{i.e., } R_{\text{max}} = [727 + s''] (\text{ft.}) \text{ lbs.}$$ E.g., if $s'' = \frac{1}{4}$ of an inch = $\frac{1}{4}$ of a foot, $R_{\text{max}} = 4589$ lbs., and the maximum value of $R$ would probably be about double this near the end of the impact. If the anvil also sinks during the impact a distance $s'''$, we must substitute $s'''' + s'''$ instead of $s''$: this will give a smaller value for $R_{\text{max}}$. 140 MECHANICS OF ENGINEERING. By mean value for $R$ is meant [eq. (c)] that value, $R_{\text{mean}}$, which satisfies the relation $$R_{\text{mean}} = \int_0^t R dt.$$ This may be called more explicitly a *space-average*, to distinguish it from a *time-average*, which might appear in some problems, viz., a value $R_{\text{mean}}$ to satisfy the relation ($\ell$ being the duration of the impact) $$R_{\text{mean}} = \int_0^\ell R dt,$$ and is different from $R_{\text{mean}}$. From $M e' = 787$ ft.-lbs., we have $e' = 19.06$ ft. per sec., whereas for a free fall it would have been $\sqrt{2} \times 32.2 \times 1 = 8.03$. (This example is virtually of the same kind as Prob. 4, §59, differing chiefly in phraseology.) 138. Pile-Driving.—The safe load to be placed upon a pile after the driving is finished is generally taken as a fraction (from $\frac{1}{4}$ to $\frac{1}{2}$) of the resistance of the earth to the passage of the pile as indicated by the effect of the last few blows of the ram, in accordance with the following approximate theory: Toward the end of the driving the resistance $R$ encountered by the pile is nearly constant, and it is assumed to be that met by the ram at the moment when the resistance $e'$ through which the head of the pile sinks as an effect of the last blow is observed. If $G$, then, is the weight of the ram, $= Mg$, and $h$ the height of free fall, the velocity due to $h$, on striking the pile, is $v = \sqrt{2gh}$ (§ 59), and we have, from eq. (c), $$\frac{1}{2} M e', \text{i.e., } Gh = \int_0^{\ell} R dt = R e'. \quad (1)$$ (R being considered constant); hence $R = Gh + e'$, and the safe load (for ordinary wooden piles), $$P = \text{from } \frac{1}{4} \text{ to } \frac{1}{2} \text{ of } Gh + e'. \quad (2)$$ Maj. Sanders recommends $\frac{1}{4}$ from experiments made at Fort A diagram showing a pile with a cross-section labeled "M" and "H". The text describes the calculation of safe load for pile-driving. **WORK, ENERGY, AND POWER.** 141 Delaware in 1851; Molesworth, J.; General Barnard, I., from extensive experiments made in Holland. Of course, by (9), given $P$, we can compute $\epsilon$'. (Owing to the uncertainty as to how much of the resistance $R$ is due to friction of the soil on the sides of the pile, and how much to the inertia of the soil around the shoe, the more elaborate theories of Weisbach and Rankine seem of little practical account.) 137. Example.--In preparing the foundation of a bridge-pier it is found that each pile (placing them 4 ft. apart) must bear safely a load of 10 tons. If the rails weigh one ton, and rails 12 ft., what would be the effect of the last blow on each pile ? Using the foot-ton-second system of units, and Molesworth factor $\frac{1}{4}$, eq. (9) gives $$\epsilon = \frac{1}{4}(1 \times 12 + 72) = \frac{84}{4} = \frac{21}{\text{ft}}$$ That is, the pile should be driven until it sinks only $\frac{2}{4}$ inch under each of the last few blows. 138. Kinetic Energy Lost in Inelastic Direct Central Impact.--Referring to § 60, and using the same notation as there given, we find that if the united kinetic energy possessed by two inelastic bodies after their impact, viz., $\frac{1}{2}M_1C^2 + \frac{1}{2}M_2C^2$, $C$ having the value $(M_1 + M_2)c + (M_1 + M_2)$, be deducted from the amount before impact, viz., $\frac{1}{2}M_1c_1^2 + \frac{1}{2}M_2c_2^2$, the loss of kinetic energy during impact of two inelastic bodies is $$W = \frac{1}{2}M_1M_2(c_1 - c_2)^2$$ (1) An equal amount of energy is also lost by partially elastic bodies during the first period of the impact, but is partly regained in the second. If the bodies were perfectly elastic, we would find it wholly regained and the resultant loss zero, from the equations of § 60 ; but this is not quite the reality, on account of internal vibrations. The kinetic energy still remaining in two inelastic bodies after impact (they move together as one mass) is 142 MECHANICS OF ENGINEERING. If $M_1 + M_2$ act, or, after inserting the value of $$C = (M_1 + M_2) + (M_1 + M_2),$$ we have $$W = \frac{1}{2} \left( M_1 + M_2 \right)^2.$$ (2) Example 1.—The weight $G = Mg$ falls freely through a height $h$, impinging upon a weight $G' = Mg$, which was initially at rest. After their (inelastic) impact they move on together with the combined kinetic energy just given by (2). Hence, since $\sqrt{gh}$ and 0, may be reduced to a simpler form. This energy is soon absorbed in overcoming the flange-pressure $R$, which is proportional (so long as the elasticity of the rod is not impaired) to the elongation $s$, as with an ordinary spring. If from previous experiment it is known that a force $R'$ produces an elongation $s'$, then the variable $R = (R' + s')e$. Neglecting the weight of the two bodies as a working force, we now have, from eq. (6), $$0 = \frac{R'}{s'} \int s'ds + 0 - \frac{M'gh}{M_1 + M_2},$$ i.e., $$\frac{R'}{s'} \cdot s' = \frac{M'gh}{M_1 + M_2}.$$ (3) When $s = s'$, i.e., when the masses are (momentarily) at rest in the lowest position, the flange-pressure or tensile stress in the rod is a maximum, $R' = (R' + s')e$, whence $e' = R'e_s + R's'$; and (3) may be written $$\frac{R'}{2} e' = \frac{M'gh}{M_1 + M_2}.$$ (4) or $$\frac{R'e_s}{2} = \frac{M'gh}{M_1 + M_2}. \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (5)$$ Eq. (8) gives the final elongation of the rod, and (5) the greatest tensile force upon it, provided the elasticity of the rod is not **WORK, ENERGY, AND POWER.** 143 impaired. The form $\frac{dR}{ds}$ in (4) may be looked upon as a direct integration of $f R ds$, viz., the mean resistance $(\frac{dR}{ds})$ multiplied by the whole distance $(s)$ gives the work done in overcoming the variable $R$ through the successive $ds$. If the elongation is considerable, the working-forces $G_s$ and $G_m$ cause a certain amount of elongation to appear in the term $(+G_s + G_m s)$ in the right-hand members of (3), (4), and (5). The upper end of the rod is firmly fixed, and the rod itself is of small mass compared with $M_s$ and $M_m$. **Example 2.—Two cars, Fig. 151, are connected by an elastic chain on a horizontal track. Velocities before impact (i.e., before the stretching of the chain become $c_0$ o $c_0$ gms, by means of which they are brought to a common velocity at the instant of impact, see Fig. 152.) The elongation $s'$ of the chain) is $c_c = c_0$, and $c_0 = 0$. During the stretching, i.e., the first period of the impact, the kinetic energy lost by the masses has been expended in stretching the chain, i.e., in doing the work $\frac{dR}{ds}$; hence we may write (the elasticity of the chain not being impaired) (see eq. (1)) $$\frac{M_s M_m s'}{2} = \frac{1}{2} R's' - \frac{R}{s'} = \frac{R's'}{2}$$ (6) in which the different symbols have the same meaning as in Example 1, in which the rod corresponds to the chain of this example. In this case the mutual accommodation of velocities is due to the presence of the chain, whose stretching corresponds to the compression (of the parts in contact) in an ordinary impact. In numerical substitution, 92.2 for $g$ requires the use of the units foot and second for space and time, while the unit of force may be anything convenient. **139. Work and Energy in Rotatory Motion. Axis Fixed.—** The rigid body being considered free, let an axis through $O$ perpendicular to the paper be the axis of rotation, and resolve all forces not intersecting the axis into components parallel 144 MECHANICS OF ENGINEERING. and perpendicular to the axis, and the latter again into components tangent and normal to the circular path of the point of application. These tangential components are evidently the only ones of the three sets mentioned which have moments about the axis, those having moments of the same sign as $\omega$ (the angular velocity at any instant) being called working forces, $T_x$, $T_y$, etc.; those of opposite sign, resistances, being denoted by $-T_x$, $-T_y$. The line $dt$ the point of application $B_0$, of $T_x$, describes the small arc $da = a\,ds$, whose projection on $T_x$ is $d\theta_x$, this projection falls ahead (i.e., in direction of force) of the position of the point at the beginning of $dt$, while the reverse is true for $T'_x$. From eq. (XIV), § 114, we have for $\theta$ (angul. accel.) $$\theta = \frac{(T_xa + T_ya' + \ldots) - (T'_xa' + T'_ya' + \ldots)}{I},$$ (1) which substituted in $\omega d\theta = \theta da$ (from § 110) gives (remembering that $da = da'$, etc.), after integration and transposition, $$\int_0^{\theta} T_xda + \int_0^{\theta} T_yda' + \text{etc.} = \int_0^{\theta} I'd\theta - \int_0^{\theta} I'd\theta',$$ (2) where $0$ and $\pi$ refer to any two (initial and final) positions of the rotating body. Eq. (4), § 120, is an example of this. Now $\dot{\varphi}_xI = \dot{\varphi}_yI'dM/dt = f(\ddot{\varphi}\omega d\theta dt)$, which, since $\omega d\theta$ is the angular displacement made by the body at any (that) instant, is nothing more than the sum of the amounts of kinetic energy possessed at this instant by all the particles of the body; a similar statement may be made for $\dot{\varphi}_yI$. Eq. (2) therefore may be put into words as follows: Between any two positions of a rigid body rotating about a fixed axis, the work done by the working forces is partly used in overcoming the resistance, and the remainder in changing the kinetic energy of the individual particles. If in any case **WORK, ENERGY, AND POWER** 145 this remainder is negative, the final kinetic energy is less than the initial, i.e., the work done by the working forces is less than that necessary to overcome the resistance through their respective spaces and the deficiency is made up by the recovery of some of the lost kinetic energy by the body. A moving fly-wheel, then, is a reservoir of kinetic energy. Eq. (9) has already been illustrated numerically in § 191, where the additional relation was utilized (for a connecting-rod and piston of small mass), that the work done in the steam-cylinder is the same as that done directly at the crank-pin by the working-force there. 140. Work of Equivalent Systems the Same.—If two plane systems of forces acting on a rigid body are equivalent (§ 15a), the aggregate work done by either of them during a given slight displacement or motion of the body parallel to their plane is the same. By aggregate work is meant what has already been defined as the sum of the “ virtual momenta” (§ 61 to 64), in any small displacement of the body, viz., the algebraic sum of the products, $\mathfrak{Z}(Pdu)$, obtained by multiplying each force by the projection $(du)$ of its point of application upon the force. (We here draw attention to this definition of working-forces.) Call the systems $A$ and $B$; then, if all the forces of $B$ were reversed in direction and applied to the body along with those of $A$, the compound system would be a balanced system, and hence we would have (§ 64), for a small motion parallel to the plane of the forces, $$\mathfrak{Z}(Pdu)=0,$$ i.e., $\mathfrak{Z}(Pdu)$ for $A-\mathfrak{Z}(Pdu)$ for $B=0,$ or $$+\mathfrak{Z}(Pdu) \text{ for } A = +\mathfrak{Z}(Pdu) \text{ for } B.$$ But $\mathfrak{Z}(Pdu)$ for $A$ is the aggregate work done by the forces of $A$ during the given motion, and $+\mathfrak{Z}(Pdu)$ for $B$ is a similar quantity for the forces of $B$ (not reversed) during the same small motion if $B$ acted alone. Hence the theorem is proved, and could easily be extended to space of three dimensions. 11 148 MECHANICS OF ENGINEERING. 141. Relation of Work and Kinetic Energy for any Extended Motion of a Rigid Body Parallel to a Plane. (If at any instant any of the forces acting are not parallel to the plane mentioned, their components lying in or parallel to that plane, will be used instead, since the other components obviously would be neither working forces nor resistances.) Fig. 153 shows an initial position, $q_0$ of the body; final, $n_1$; and any intermediate, $n$. The forces of the system acting may vary in any manner during the motion. In this motion each $dM$ describes a curve of its own with varying velocity $v$, tangential acceleration $p_n$ and radius of curvature $r$; hence in any position $q$, an imaginary system $B$ (see Fig. 154), equivalent to the actual system $A$ (at $g$ in Fig. 153), would be formed by applying to each $dM$ a tangential force $dT = dMp_n$, and a normal force $dN = -dMv^2/r$. In successive small displacements, the body passes from $o$ to $n$. In the small displacement of which $g$ is the initial position, each $dM$ describes a space $ds$, and $dT$ does the work $dTds = dMvds$, while $dNds$ does the work $dN \times 0 = 0$. Hence the total work done by $B$ in the small displacement at $q$ would be $$= dM'v'dv' + dM'e'dv' + \text{etc.,} \quad \ldots \quad (1)$$ including all the $dM'$s of the body and their respective velocities at this instant. But the work at $q$ in Fig. 153 by the actual forces (i.e., of system $A$) during the same small displacement must (by § 140) be equal to that done by $B$, hence $$P_{dn} + P_{dn}, + \text{etc.} = dM'v'dv' + dM'e'dv' + \text{etc.} \quad (g)$$ Now conceive an equation like $(g)$ written out for each of WORK, ENERGY, AND POWER. 147 the small consecutive displacements between positions $o$ and $n$ and corresponding terms to be added; this will give $$\int P_d du + \int P_d dv + \text{etc.}$$ $$= dM \int v^* dv' + dM \int v'' dv'' + \text{etc.}$$ $$= dM (v_n - v_n) + dM (v_n - v_n) + \text{etc.}$$ The second member may be rewritten so as to give, finally, $$\int P_d du + \int P_d dv + \text{etc.} = Z(dM v_n) - Z(dM v_n), (X.V.)$$ or, in words, the work done by the acting forces (treating a resistance as a negative working force) between any two positions is equal to the gain (or loss) in the aggregate kinetic energy of the particles of the body between the two positions. To avoid confusion, $Z$ has been used instead of the sign $f$ in one member of $(X.V.)$, in which $v_n$ is the final velocity of any $dM$ (not the same for all necessarily) and $v_0$ the initial. (The same method of proof can be extended to three dimensions.) Since kinetic energy is always essentially positive, if an expression for it comes out negative as the solution of a problem, some impossible conditions have been imposed. 142. Work and Kinetic Energy in a Moving Machine.—Defining a mechanism or machine as a series of rigid bodies joined or connected together, so that working forces applied to one or more may be the means of overcoming resistances occurring at other points, we see that the total amount of kinetic energy of the moving masses, let us for simplicity consider a machine the motions of whose parts are all parallel to a plane, and let all the forces acting on any one piece, considered free, at any instant be parallel to the same plane. Now consider each piece of the machine, or of any series of its pieces, as a free body, and write out eq. $(X.V.)$ for it between any two positions (whatever initial and final positions are) A diagram showing a simple machine with various components labeled. 148 MECHANICS OF ENGINEERING. selected for the first piece, those of the others must be corresponding initial and corresponding final positions), and it will be found, on adding up corresponding members of these equations, that the terms involving those components of the mutual pressures which are tangential to the rubbing surfaces will cancel out, while their components tangential to the rubbing surfaces (i.e., friction, since if the surfaces are perfectly smooth there can be no tangential action) will appear in the algebraic addition as resistances multiplied by the distances rubbed through, measured on the rubbing surfaces. For example, Fig. 155, where one rotating piece both presses and rubs on another. Let the normal pressure between them at $A$ be $P_1 = P_2$; it is a working-force for $M$, but a resistance for $M'$, hence the separate symbols for the numerically equal forces (action and reaction). Similarly, the friction at $A$ is $R_1 = R_2$; a resistance for $M'$, a working-force for $M''$. (In some cases, of course, friction may be a resistance for both bodies.) For a small motion, $A$ describes the small arc $AA'$ about $O'$ in dealing with $M$, but for $M''$ it describes the arc $AA''$ about $O''$, $AA''$ being parallel to the surface of contact $AD$, while $AB$ is perpen- Fig. 155. dicular to $AA'$. In Figs. 156 and 157 we see $M'$ and $M''$ free, and their corresponding small rotations indicated. During these motions the kinetic energy (K.E.) of each mass has changed by amounts $\delta(K.\mathrm{E})_{AB}$ and $\delta(K.\mathrm{E})_{AC}$ respectively, and hence eq. (XV) gives, for each free body in turn, $$P_1 \cdot u_1 - R_1 \cdot AB - R_2 \cdot AB = d(K.\mathrm{E})_{AB} \quad \text{(1)}$$ $$P_2 \cdot u_2 - R_2 \cdot AB + R_1 \cdot AB = d(K.\mathrm{E})_{AC} \quad \text{(2)}$$ W O R K , E N E R G Y , A N D P O W E R 149 Now add (1) and (2), member to member, remembering that $P_i = R_i$ and $P_i = R_i = F_{\text{friction}}$, and we have $$P_{i} \text{as} - F_i A^{\prime}A^{\prime} - R_i \text{as} = d(K.E.)_i + d(K.E.)_j, \quad (3)$$ in which the mutual actions of $M'$ and $M''$ do not appear, except the friction, the work done in overcoming which, when the two bodies are thus considered collectively, is the product of the friction by the distance $A^{\prime}A^{\prime}$ of actual rubbing measured on the rubbing surface. For any number of pieces, then, whatever may be the number of motions at any instant of this article is true, since any finite motion consists of an infinite number of small motions to each one of which equation like (8) is applicable. Summing the corresponding terms of all such equations, we have $$\int P_i du + \int P_j du + \text{etc.} = Z(K.E.)_i - Z(K.E.)_j \quad (XVI)$$ This is of the same form as (XV), but instead of applying to a single rigid body, deals with any assembly of rigid parts forming a machine, or any part of a machine (a similar proof will apply to different kinds of space); but it must be remembered that it excludes all the mutual actions of the pieces considered except friction, which is to be introduced in the manner just illustrated. A flexible inextensible cord may be considered as made up of a great number of short rigid bodies joined without friction, and hence may form part of a machine without vitating the truth of (XVI). $Z(K.E.)_i$ signifies the sum obtained by adding the amounts of kinetic energy $(\frac{1}{2}mv^2)$ for all elementary masses possessed by all the parts of a rigid body at their final positions; $Z(K.E.)_j$, a similar sum at their initial positions. For example, the K.E. of a rigid body having a motion of translation of velocity $v = \frac{1}{2}mv^2 = \frac{1}{2}mv^2$; that of a rigid body having an angular velocity $\omega$ about a fixed axis $Z = \frac{1}{2}\omega^2I$ ($§ 130$); while, if it has an angular velocity $\omega$ about a gravity- 150 MECHANICS OF ENGINEERING. axis $Z_0$ which has a velocity $v_{z0}$ of translation at right angles to itself, the (K. E.) at this instant may be proved to be $$\frac{1}{2}Mv_{z0}^2 + \frac{1}{2}\omega^2I_n,$$ i.e., is the sum of the amounts due to the two motions separately. 143. K. E. of Combined Rotation and Translation.—The last statement may be thus proved. Fig. 158. At a given instant the velocity of any $dM$ is $v$, the diagonal formed on the velocity $v_z$ of translation, and the rotary velocity $\omega z$ relatively to the moving gravity-axis $Z$ (perpendicular to paper) (see § 71), therefore, $$v = v_z^2 + (\omega z)^2 - 2(\omega z)v_z\cos\varphi;$$ hence we have K. E. at this instant, $$= f'dMv^2 = f'dMv_z^2 + 4\omega^2f'dM\omega z - 4\omega^2f'dM\omega z\cos\varphi,$$ but $\rho\cos\varphi = y$, and $fdMy = Mg = 0$, since $Z$ is a gravity-axis, $$\therefore K.E. = \frac{1}{2}Mv_z^2 + \frac{1}{2}\omega^2I_n.$$ Q. E. D. It is interesting to notice that the K. E. due to rotation, viz., $\frac{1}{2}\omega^2I_n = \frac{1}{2}M(\omega b)^2$, is the same as if the whole mass were concentrated in a point, line, or thin shell, at a distance $k$, the radius of gyration, from the axis. 144. Example of a Machine in Operation.—Fig. 159. Consider the four consecutive moving masses, $M'$, $M''$, $M'''$, and $M''''$ (being the piston; connecting-rod; fly-wheel; crank, drum, and chain; and weight on inclined plane) as free, collectively. Let us apply eq. (XV.1), the initial and final positions being taken when the crank-pin is at its dead-points and $a$; i.e., we deal with the progress of the pieces made while the crank-pin describes its upper semicircle. Remembering that the mutual actions between any two of these four masses can be left out of account (except friction), the only forces to be put in are the actions of other bodies on each one of these four, and are A diagram showing a crank mechanism with a piston, connecting rod, flywheel, crank, drum, chain, and weight on an inclined plane. WORK, ENERGY, AND POWER. 151 shown in the figure. The only mutual friction considered will be at the crank-pin, and if this as an average $F^{\prime}$, the work done on it between $o$ and $n = h^{\prime} \pi r$, where $r^{\prime}$ = radius of crank-pin. The work done by $P$, the effective steam-pressure (let it be constant) during this period is $F^{\prime} t$; that done upon the weight $G^{\prime}$ of connecting-rod is cancelled by the work done by $\mathscr{H}$ in the descent following; the work done upon $G^{\prime} = G^{\prime} \pi a \sin \beta$, where $a =$ radius of drum; that upon the friction $F_{\mathrm{fr}} = F_{\mathrm{fr}} \pi a$. The pressures $N$, $N^{\prime}$, $N^{\prime\prime}$, and weights $G$ and $G^{\prime\prime}$ are neutral, i.e., do no work either positive or negative. Hence the left-hand member of (XVI.) becomes, between $o$ and $n$, $$P.F - F^{\prime}.F - F^{\prime\prime}.F - G^{\prime}\pi a \sin \beta - F_{\mathrm{fr}} \pi a,\quad(1)$$ provided the respective distances are actually described by these forces, i.e., if the masses have sufficient initial kinetic energy to carry the crank-pin beyond the point of minimum velocity, with the aid of the working force $P$, whose effect is small up to that instant. As for the total initial kinetic energy, i.e., $\mathscr{S}(K.E.)$, let us express it in terms of the velocity of the crank-pin at $o$, viz., $V_0$. The $(K.E.)$ of $M$ is nothing ; that of $M^{\prime}$ which at this instant is equal to zero. The angular velocity (found for the instant) with angular velocity $\omega' = V_0 + V'$, is $\omega'' = V_0 + V''$, that of $M''$ is $\omega''' = V_0 + r$, in which $\omega''' = V_0 + r$; that of $M^{(\mathrm{trans})} = M^{(\mathrm{rev})}$, in which $v^{(\mathrm{rev})} = (a + r)V_0$. $\mathscr{S}(K.E.)$ is expressed Fig. 138. 151 152 MECHANICS OF ENGINEERING. in a corresponding manner with $V_s$ (final velocity of crank-pin) instead of $V_r$. Hence the right-hand member of (XVI.) will give putting the radius of gyration of $M'$ about $O' = k'$, and that of $M''$ about $C = k$, $$\left(V_s - V_r \right)\left[\frac{M'}{r_1} + \frac{M''}{r_2}\right] = 0.$$ (2) By writing (1) = (2), we have an equation of condition, capable of solution for any one unknown quantity, to be satisfied for the extent of motion considered. It is understood that the chain is always taut, and that its weight and mass are neglected. 145. Numerical Case of the Forgoing.---(Foot-pound-second system of units for space, force, and time; this requires $g = 32.2$.) Suppose the following data :
Pear. Lbs. Lbs. Mass Units
$p = 3.0$ $P_1 = 6000$ $G' = 60$ 1.86
$p' = 4.0$ $F_1 = 200$ $G'' = 50$ $H' = 1.35$
$a = 5.0$ $F'' (w.r.t)$ $400$ $G'' = 12.4$
$r = 1.0$ $F_2 = 800$ $G'' = 8250$ $H'' = 100.9$
$k' = 1.8$
$k'' = 0.1$
Also let $V_s = v$, per sec., $\beta = 30^\circ$ Denote (1) by $W$ and the large bracket in (2) by $M$ (this by some is called the total mass "reduced" to the crank-pin). Putting (1) = (2) we have, solving for the unknown $V_s$ $$V_s = \sqrt{\frac{W^2}{M}} + V_r.$$ (3) For above values, $$W = 12,900 - 400 - 125.7 - 7590.0 - 1417.3$$ $$= 2467 \text{ foot-pounds};$$ while $$M = 0.5 + 40.3 + 295.0 = 365.8 \text{ mass-units};$$ whence $$V_s = \sqrt{18.56 + 16} = \sqrt{34.56} = 5.88 \text{ ft. per second}.$$ WORK, ENERGY, AND POWER. 153 As to whether the crank-pin actually reaches the dead-point $n$, requires separate investigations to see whether $V$ becomes zero or negative between $o$ and $n$ (a negative value is indmissible, since a reversal of direction implies a different value for $V$), i.e., whether the proposed extent of motion is realised ; and these are met by assigning some other intermediate position $m$, a final position being thus obtained, remem- bering that when $m$ is not a dead-point the $(K.E.)_m$ of $M$ is not zero, and must be expressed in terms of $\Gamma_m$, and that the $(K.E.)_m$ of the connecting-rod $M'$ must be obtained from § 143. 146. Regulation of Machines.—As already illustrated in several examples (§ 121), a fly-wheel of sufficient weight and radius may prevent too great fluctuation of speed in a single stroke of an engine; but to prevent a permanent change, which must occur if the work of the working force or forces (such as the steam-pressure on a piston, or water-impulse in a turbine) exceeds for several successive strokes or revolutions the work required for overcoming friction (including also the re- istance at the teeth of saws, etc., etc.) through their respective spaces, automatic governors are employed to diminish the working force, or the distance through which it acts per stroke, until the normal speed is restored; or vice versa, if the speed slackens, as when new resistances are temporarily brought into play. Hence when several successive periods, strokes (or other cycles) are considered, the kinetic energy of the moving parts will disappear from eq. (XVI), leaving it in this form: work of working force = work done upon resistances. 147. Power of Motors.—In a mill where the same number of machines are run continuously at a constant speed proper for their work, turning out per hour the same number of barrels of flour, feet of lumber, or other commodity, the motor (e.g., a steam-engine, or turbine) works at a constant rate, i.e., de- velops a definite horse-power (H.P.), which is thus found in the case of steam-engines (double-action):
WORK, ENERGY, AND POWER.153
As to whether the crank-pin actually reaches the dead-point $n$, requires separate investigations to see whether $V$ becomes zero or negative between $o$ and $n$ (a negative value is indmissible, since a reversal of direction implies a different value for $V$), i.e., whether the proposed extent of motion is realised ; and these are met by assigning some other intermediate position $m$, a final position being thus obtained, remem-
remembering that when $m$ is not a dead-point the $(K.E.)_m$ of $M$ is not zero, and must be expressed in terms of $\Gamma_m$, and that the $(K.E.)_m$ of the connecting-rod $M'$ must be obtained from § 143.
146. Regulation of Machines.—As already illustrated in several examples (§ 121), a fly-wheel of sufficient weight and radius may prevent too great fluctuation of speed in a single stroke of an engine; but to prevent a permanent change, which must occur if the work of the working force or forces (such as the steam-pressure on a piston, or water-impulse in a turbine) exceeds for several successive strokes or revolutions the work required for overcoming friction (including also the re-
istance at the teeth of saws, etc., etc.) through their respective spaces, automatic governors are employed to diminish the working force, or the distance through which it acts per stroke, until the normal speed is restored; or vice versa, if the speed slackens, as when new resistances are temporarily brought into play. Hence when several successive periods, strokes (or other cycles) are considered, the kinetic energy of the moving parts will disappear from eq. (XVI), leaving it in this form:
work of working force = work done upon resistances.
147. Power of Motors.—In a mill where the same number of machines are run continuously at a constant speed proper for their work, turning out per hour the same number of barrels of flour, feet of lumber, or other commodity, the motor (e.g., a steam-engine, or turbine) works at a constant rate, i.e., de-
velops a definite horse-power (H.P.), which is thus found in
the case of steam-engines (double-action):
154 MECHANICS OF ENGINEERING. H.P. = total mean effective steam-pressure on piston in lbs. $\times$ distance by pie-travelled per second + 550, ton per second i.e., the work (in ft.-lbs.) done per second by the working force divided by 550 (see § 132). The total effective pressure at any instant is the excess of the forward over the back-pressure, and by its mean value (since steam is usually used expansively) is meant such a value $P'$ as, multiplied by the length of stroke $l$, shall give $$P't = \int P'dz,$$ where $P$ is the variable effective pressure and $dz$ an element of its path. If $u$ is the number of strokes per second, we may also write (foot-pound-second system) $$H.P. = P'lu + 550 = \left[ \frac{P'}{u} \int P'dz \right] u + 550.$$ (XVII.) Very often the number of revolutions per minute, $m$, of the crank is given, and then $$H.P. = P'(lbs.) \times 2\pi(f) \times m + 33,000.$$ If $F =$ area of piston we may also write $P' = Fp'$, where $p'$ is the mean effective steam-pressure per unit of area. Evidently, to obtain $P'$ in lbs., we multiply $F$ in lbs. by $p'$ in lbs. per sq. in., or $F$ in sq. ft. by $p'$ in lbs. per sq. foot ; the former is customary. $p'$ in practice is obtained by measurements and computations from "indicator-cards" (see § 135, in which $(P' - R')$ corresponds to $P$ of this section); or $P't$, i.e., $\int P'dz$, may be computed theoretically as in § 59, Problem 4. The power as thus found is expended in overcoming the friction of all moving parts (which is sometimes a large item), and the resistance peculiar to the kind of work done by the machines. The work periodically stored in the increasing kinetic energy of the moving masses is restored as they periodically resume their minimum velocities. WORK, ENERGY, AND POWER. 103 148. Potential Energy.—There are other ways in which work or energy is stored and then restored, as follows: First. In raising a weight $G$ through a height $h$, an amount of work = $Gh$ is done upon $G$, as a resistance, and if at any subsequent time the weight is allowed to descend through the same vertical distance $A$ (the form of path is of no account), $G$, now a working force, does the work $Oh$, and thus in aiding the motor repays, or restores, the $Gh$ expended by the motor in raising it. As $A$ is a small part of the total distance, the forces of gravity rises and sinks periodically in the motion of the machine, the force $G$ may be left out of account in reckoning the expenditure of the motor's work, and the body when at its highest point is said to possess an amount $Gh$ of potential energy, i.e., energy of position, since it is capable of doing the work $Gh$ in sinking through its vertical range of motion. Second. So far, all bodies considered have been by express stipulation rigid, i.e., incapable of changing shape. To see the effect of this lack of elasticity as affecting the principle of work and energy in machines take the simple case in Fig. 100. A helical spring at a given instant is acted on at each end by a force $P$ in an axial direction (they are equal, supposing the mass of the spring small). As the machine operates of which it is a member it moves to a new consecutive position $B$, suffering an extension or elongation $dA$ in its length (if $P$ is increasing), $P$ on the right, a working force, does the work $PdA$; how is this expended? $P$ on the left has the work $PdA$ done upon it, and the mass is too small to absorb kinetic energy or to bring its weight into consideration. The remainder, $PdA - PdA$, is expended in stretching the spring an additional amount $dA$, and is capable of restoration if the spring retains its elasticity. Hence the work done in changing the form of bodies if they are elastic is seen to restore the form of bodies again. That is, in the operation of machines, the name potential energy is also given to the energy A diagram showing a helical spring with two points labeled A and B. Point A shows a smaller extension than point B. Fig. 100 156 MECHANICS OF ENGINEERING. stored and restored periodically in the changing and regaining form of elastic bodies. **149. Other Forms of Energy.—Numerous experiments with various kinds of apparatus have proved that for every 772 (about) ft.-lbs. of work spent in overcoming friction, one British unit of heat is produced (viz., the amount of heat necessary to raise the temperature of one pound of water from 32° to 33° Fahrenheit); while from converse experiments, in which the amount of heat used in operating a steam-engine was all carefully estimated, the disappearance of a certain portion of it could only be accounted for by assuming that it had been converted into work at the same rate of (about) 772 ft.-lbs. of work to each unit of heat (or 425 kilometres to each French unit of heat). This number, 772, or 425, according to the system of units employed, is known as the "Law of Conservation of Heat," first discovered by Joule and confirmed by Hertz.* Heat then is energy, and is supposed to be of the kinetic form due to the rapid motion or vibration of the molecules of a substance. A similar agitation among the molecules of the (hypothetical) ether diffused through space is supposed to produce the phenomena of light, electricity, and magnetism. Chemical action being also considered a method of transforming energy (the possible future occurrence as in the case of coal and oil), we have thus a potent source of well-known doctrine of the Conservation of Energy, in accordance with which energy is indestructible, and the doing of work is simply the conversion of one or more kinds of energy into equivalent amounts of others, is now one of the accepted hypotheses of physics. Work consumed in friction, though practically lost, still remains in the universe as heat, electricity, or some other subtle form of energy. **150. Power Required for Individual Machines. Dynamometers of Transmission.—If a machine is driven by an endless belt from the main-shaft, A, Fig. 161, being the driving-pulley * Prof. Rowland's recent experiments result in the value 428.8 kilometres at a temperature of 5° Cent.** WORK, ENERGY, AND POWER. 157 on the machine, the working force which drives the machine, in other words the "grip" with which the belt takes hold of the pulley tangentially, $$P = P' - P''$$ and $P'$ being the tensions in the "pulling" and "following" sides of the belt respectively. The belt is supposed not to slip on the pulley. If $v$ is the velocity of the pulley-circumference, the work expended on the machine per second, i.e., the power, $$L = (P - P')v$$ (1) To measure the force $(P - P')$, an apparatus called a Dynamometer of Transmission may be placed between the main shaft and the machine, and the belt made to pass through it in such a way as to measure the tensions $P$ and $P'$, and principally their difference, without meeting any resistance in so doing; that is, the power is transmitted, not absorbed, by the apparatus. One invention for this purpose (mentioned in the Journal of the Franklin Institute some years ago) is shown (in principle) in Fig. 162. A vertical plate carrying four pulleys and a scale-pan is first balanced on the right side of the plate, then adjusted, as shown, and the power turned on, a sufficient weight $G$ is placed in the scale-pan to balance the plate again, for whose equilibrium we must have $Gb = Pa - P'a$, since the $P$ and $P'$ on the right are purposely given no leverage about $C$. The velocity of belt, $v$, is obtained by a simple counting device. Hence $(P - P')$ and $v$ become known, and : $L$ from (1). Many other forms of dynamometers have been invented for use, some applicable whether the machine is driven by belting or gearing from the main shaft. Emerson's Hydrodynamics describes his own invention on p. 283, and gives results of measurements with it; e.g., at Lowell, Mass., the power required to drive 112 looms, weaving 36-inch sections, No. 20 yarns. Fig. 161. Fig. 162. 158 MECHANICS OF ENGINEERING. 60 threads to the inch, speed 130 picks to the minute, was found to be 16 H.P., i.e., 4 H.P. to each boom (p. 838). 151. Dynamometers of Absorption.--These are so named since they furnish in themselves the resistance (friction or a weight) in the overcoming (or raising) of which the power is expended or absorbed. Of these the Parny Friction Brake is the most common, and is used for measuring the power developed by a given motor (e.g., a steam-engine or turbine) not absorbed in the friction of the motor itself. Fig. 163 A diagram showing a dynamometer of absorption. A vertical arm with a scale pan at the end is connected to a horizontal arm via a pulley system. The arm is balanced by a weight G placed on the scale pan. Fig. 163 shows one fitted to a vertical pulley driven by the motor. By tightening the bolt B, the velocity $v$ of pulley-rim may be made constant at any desired value (within certain limits) by the consequent friction. $v$ is measured by a counting apparatus, while the friction (or tangential components of action between pulley and brake), $F$, becomes known by noting the weight $G$ which must be placed in the scale pan to balance the arm between the cheeks ; then $$Fa = Gb \quad \ldots \ldots \ldots \quad (1)$$ for the equilibrium of the brake (imposing the weight of brake and scale-pan previously balanced on $G$) and the work done per unit of time, or power, is $$L = Fe \quad \ldots \ldots \ldots \quad (2)$$ A "dash-pot" is frequently connected with the arm to prevent sudden oscillations. In case the pulley is horizontal, a bellcrank lever is added between the arm and the scale-pan, and then eq. (1) will contain two additional lever-arms. A "dash-pot" is frequently connected with the arm to prevent sudden oscillations. In case the pulley is horizontal, a bellcrank lever is added between the arm and the scale-pan, and then eq. (1) will contain two additional lever-arms. **WORK, ENERGY, AND POWER.** 159 153. The Indicator, used with steam and other fluid engines, is a special kind of dynamometer in which the automatic motion of a pencil describes a curve on paper whose ordinates are proportional to the fluid pressures exerted in the cylinder at successive points of the stroke. Thus, Fig. 164, the back-pressure being constant and $=P_0$, the ordinates $P_0$, $P_1$, $P_2$, represent the effective pressures at equally spaced points of division. The mean effective pressure $P$ (see § 147) is, for this figure, by Simpson's Rule (six equal spaces). $$P = \frac{1}{6}(P_0 + 4(P_1 + P_2 + P_3) + 2(P_4 + P_5) + P_6).$$ This gives a near approximation. The power is now found by § 147. 154. The theory of Atwood's Machine is most directly expressed by the principle of work and energy; i.e., by eq. (XVI), § 142. Fig. 165. The parts considered free, collectively, are the rigid bodies $P$, $Q$, $G$, and four friction-wheels like $G_i$ and the flexible cord, which does not slip on the upper pulley. There is no slipping at $D$, hence no sliding friction there. The actions between these bodies are the resultant of the working force $P$, the resistances $Q$ and the four $F_i$'s (at bearings of friction-wheel axles); all others ($G$, $4G_i$, and the four $R_i$'s) are neutral. Since there is no rubbing between any two of the eight bodies, no mutual actions whatever will enter the equation. Let $P > Q$, and $I$ and $\bar{I}$ be the moments of inertia of $Q$ and $G_i$, respectively, about their respective axes of figure. Let the apparatus start from rest, then when $P$ has descended through any vertical distance $x$, and so— A diagram showing a pulley system with weights and pulleys labeled P, Q, G, G1, G2, G3, G4, R1, R2, R3, R4. Fig. 165 160 MECHANICS OF ENGINEERING. quired the velocity $v$, $Q$ has been drawn up an equal distance and acquired the same velocity, while the pulley $G$ has ac- quired an angular velocity $\omega = v + s$, so that friction-pull on an angle of inclination $\alpha$ has done the work $P_s Q$ upon it, while $F^2$ has done the work $P_s Q$ had the work $Qs$ done upon it, while each $F^2$ has been overcome through the space $(r_1 : a) \times (r_2 : a) x$: all the other forces are neutral. Hence, from eq. (XVI), § 142 (see also § 139), we have $$P_s - Qs - 4 F \frac{r_1}{a} \frac{r_2}{a} = P \frac{r_1}{g} \frac{w^2}{g} + \frac{1}{g} \frac{w^2}{g} + \frac{1}{g} \frac{w^2}{g} w^2 I - 0.$$ Evidently $v = \sqrt{s} \times constant$, i.e., the motion of $P$ and $Q$ is uniformly accelerated. If, after the observed space $s$ has been described, $P$ is suddenly diminished to such a value $P'$ that the motion continues with a constant velocity $v$, we shall have, for any further space $s'$, $$P' s' - Qs' - 4 F \frac{r_1}{a} \frac{r_2}{a} = 0,$$ from which $F$ can be obtained (nearly); while if $t'$ be the ob- served time of describing $s'$, $v = s' + t'$ becomes known. Also we may write $I = (G + g) k^2$ and $I_c = (G_c + g_c) k^2$, and thus finally compute the acceleration of gravity, $g$, from our first equation above. 154. Boat-Rowing—fig. 168. During the stroke proper, let $P_s$ be pressure between oar-handle; hence the pressures on the foot-rest are $2P_r$ resistances. Let $M =$ mass of boat and load, $v_c$ and $v_a$ its velocities at beginning and end of stroke. $P_s =$ pressures between oar-blade and water. $R =$ mean re- istance of water to the boat's passage at this (mean) speed. These are the only (horizontal) forces to be considered as act- ing on the boat and two oars, considered free collectively. During the stroke the boat describes the space $s = CD$, the oar-handle the space $s_1 = AB$, while the oar-blade slips back- A diagram showing a boat being rowed by two oars. The diagram includes labels for various parts of the boat and oars. WORK, ENERGY, AND POWER. 161 ward through the small space (the "slip") = s, (average). Hence by eq. (XVI), § 149, $$2P_{a}-2P_{b}-R_{a}-2P_{c}=4M(v_{a}-v_{b});$$ i.e., $2P(s-a)=2P\times AE=2P\times R_{a}+2P\times v_{a}+\frac{1}{2}M(v_{a}^{2}-v_{b}^{2})$; or, in words, the product of the car-hullside pressures into the distance described by them measured on the boat, i.e., the work done by these pressures relatively to the boat, is entirely accounted for in the work of slip and of liquid-resistance, and increasing the kinetic energy of the mass. (The useless work due to slip is inevitable in all paddle or screw propulsion, as well as in any amount lost in machine-friction, not considered in the present problem.) During the "recover" the velocity decreases again to zero. 155. Examples.-1. What work is done on a level track, in bringing up the velocity of a train weighing 300 tons, from zero to 30 miles per hour; if the total frictional resistance (at any velocity, say) is 10 lbs. per ton and if this change of speed is accomplished in a length of 3000 feet? (Foot-ton-second system.) 30 miles per hour = 44 ft. per sec. The mass $$=300+32.2=6.2;$$ ∴ the change in kinetic energy, $$=(\frac{1}{2}Mv^{2}-\frac{1}{2}M\times0),$$ $$=\frac{1}{2}(6.2)\times44^{2}=6001.6\;\mathrm{ft.\cdot}\mathrm{tons}.$$ 162 MECHANICS OF ENGINEERING. The work done in overcoming friction is $F_s$, i.e., $$= 200 \times 10 \times 3000 = 6,000,000 \text{ ft.-lbs.} = 3000 \text{ ft.-tons};$$ :. total work $$= 6001.6 + 3000 = 9001.6 \text{ ft.-tons}.$$ (If the track were an up-grade, 1 in 100 say, the item of 300 × 30 = 90,000 ft.-tons would be added.) Example 1.—Required the rate of work, or power, in Example 2. The power varies with the velocity of the train at any instant. Assume the motion to be uniformly accelerated, then the working force is constant; call it $P$. The acceleration ($\frac{5}{6}$) will be $p = v^2 - 2a = 198.6 - 6000 = 0.392$ ft. per sec., and since $P = F = Mw$, we have $$P = 1 \text{ ton} + (30 + 32) \times 0.392 = 3 \text{ tons},$$ which is $6000 + 300 = 30$ lbs. per ton of train, of which 20 is due to its inertia, since when the speed becomes uniform the work of the engine is expended on friction alone. Hence when the velocity is $44$ ft. per sec., the engine is working at the rate of $Pv = 264,000$ ft.-lbs. per sec., i.e., at the rate of $480$ H. P. At $44$ ft. $= 6,000$ ft. from the start, at the rate of $340$ H. P., half as much power would be used. At a uniform speed of $30$ miles an hour the power would be simply $1 \times 44 = 44$ ft.-tons per sec. = $160$ H. P. Example 3.—The resistance offered by still water to the passage of a certain steamer at $10$ knots an hour is $15,000$ lbs. What power must be developed by its engines, at this uniform speed, considering no loss in "slip" nor in friction of machinery? Ans. 461 H. P. Example 4.—Same as 3, except that the speed is $15$ knots (i.e., nautical miles; each = $698$ feet) an hour, assuming that the resistances are as the square of the speed (approximately true). Ans. 1556 H. P. Example 5.—Same as 3, except that $12\%$ of the power is absorbed in the "slip" (i.e., in pushing aside and backwards the water acted on by the screw or paddle), and $5\%$ in friction of machinery. Ans. 576 H. P. Example 6.—In Example 3, if the crank-shaft makes $60$ A diagram showing a steam engine with pistons and connecting rods. WORK, ENERGY, AND POWER. 163 revolutions per minute, the crank-pin describing a circle of 18 inches radius, required the average* value of the tangential component of the thrust (or pull) of the connecting-rod against the crank-pin. **Example 7** —A solid sphere of cast-iron is rolling up an incline of $30^\circ$, and at a certain instant its centre has a velocity of 36 inches per second. Neglecting friction of all kinds, how much further will the ball mount the incline (see p. 158)? $$Ans.\ 10890\ \mathrm{lbs}.$$ **Example 8** —In Fig. 163, with $b = 4$ ft. and $a = 16$ inches, it is found in one experiment that the friction which keeps the speed of the pulley at 120 revolutions per minute is balanced by a weight $G = 160$ lbs. Required the power thus measured. $$Ans.\ 14.6\ \mathrm{H.P.}$$ Although in Examples 1 to 6 the steam cylinder is itself in motion, the work per stroke is still = mean effective steam-pressure $\times$ piston $\times$ length of stroke, for this is the final form to which the total work must be reduced. Work done by, or upon, the two cylinder heads and the two sides of the piston will reduce, when added algebraically. See § 154. * By "average value" is meant such a value, $T_n$, as multiplied into the length of path described by the crank-pin per unit of time shall give the power exerted. A diagram showing a crank mechanism with a pulley and weights. 164 MECHANICS OF ENGINEERING. CHAPTER VII FRICTION. 158. Sliding Friction.--When the surfaces of contact of two bodies are perfectly smooth, the direction of the pressure or pair of forces between them is normal to these surfaces, i.e., to their tangent-plane; but when they are rough, and one moving one on the other, the force or forces acting between them are away from the normal, each on the side opposite to the direction of the (relative) motion of the body on which it acts. Thus, Fig. 167, a block whose weight is $G$, is drawn on a rough horizontal table by a horizontal cord, the tension in which is $P$. On account of the roughness of one or both bodies the action of the table upon the block is a force $P$, inclined to the normal (or) vertical $N$ at an angle $\varphi$, away from the direction of the relative velocity $v$. This angle $\varphi$ is called the angle of friction, while the tangential component of $P$, which in this case is equal and opposite to $G$ the weight of the body, is called the normal pressure. Obviously $F = N \tan \varphi$, and denoting tan $\varphi$ by $f$, we have $$F = fN.$$ (1) $f$ is called the coefficient of friction, and may also be defined as the ratio of the friction $F$ to the normal pressure $N$ which produces it. **FRICITION.** 165 In Fig. 167, if the motion is accelerated (sec. $p$), we have (eq. (1.V.), § 38) $P - F = Mg$, if uniform, $F = F_0$; from which equations (see also (1)) $F$ may be computed. In the latter case, the surfaces are in contact with different velocities (the surfaces retaining the same velocity of course); then a uniformly accelerated motion is impossible unless $P - F$ were constant. As for the lower block or table, forces the equals and opposites of $N$ and $F$ (or a single force equal and opposite to $P_1$) are comprised in the surfaces acting upon it. As to whether $F$ is a working force or a resistance, when either of the two bodies is considered free, depends on the circumstances of its motion. For example, in friction-gearing the tangential action between the two pulleys is a resistance for one, a working force for the other. If the force $P$, Fig. 167, is just sufficient to start the body, or is just on the point of starting it (this will be called impending motion), $F$ is called the friction of rest. If the body is at rest and $P$ is not sufficient to start it, the tangential component will then be < $f$ the friction of rest, viz., just = $P$. As $P$ increases, this component continually equals it in value, and $P$, acquires a direct component, inclined from the horizontal, until the instant of impending motion; when the tangential component is = $fX$ = the friction of rest. When motion is once in progress, the friction, called then the friction of motion, = $fX$, in which $f$ is not necessarily the same as in the friction of rest. 157. Laws of Sliding Friction.—Experiment has demonstrated the following relations approximately, for two given rubbing surfaces : (See § 173.) (1) The coefficient, $f$, is independent of the normal pressure $N$. (2) The coefficient, $f$, for friction of motion, is the same at all velocities. (3) The coefficient, $f$, for friction of rest (i.e., impending motion) is usually greater than that for friction of motion (probably on account of adhesion). A diagram showing two blocks on a table with a force applied to one block. 166 MECHANICS OF ENGINEERING. (4) The coefficient, $f$, is independent of the extent of rubbing surface. (5) The interposition of an unguent (such as oil, lard, tallo w, etc.) diminishes the friction very considerably. 168. Experiments on Sliding Friction.—These may be made with simple apparatus. If a block of weight = $G$, Fig. 168, be placed on an inclined plane of uniformly rough surface, and the latter be gradually more and more inclined from the horizontal until the block begins to move, the value of $\beta$ at this instant = $\varphi$, and tan $\varphi = f =$ coefficient of friction of rest. For from $S X = 0$ we have $F$, i.e., $fN = G \sin \beta$; from $S Y = 0$, $N = G \cos \beta$; whence tan $\beta = f$; : $\beta$ must = $\varphi$. Suppose $\beta$ so great that the motion is accelerated, the body starting from rest at $o$, Fig. 169. It will be found that the distance $x$ varies as the square of the time, hence (§ 56) the motion is uniformly accelerated (along the axis $X$). (Notice in the figure that $G$ is no longer equal and opposite to $P_x$, the resultant of $N$ and $P_y$ as in Fig. 168.) $$S Y = 0,$$ which gives $N - G \cos \beta = 0;$ $$S X = Mp,$$ which gives $G \sin \beta - f N = (G + g)p;$ while (from § 56) $$p = 2x + e.$$ Hence, by elimination, $x$ and the corresponding time $t$ having been observed, we have for the coefficient of friction of motion $$f = \tan \beta - \frac{2e}{gt^2 \cos \beta}.$$ **FRICTION.** 167 In view of (8), § 157, it is evident that if a value $\beta_m$ has been found experimentally for $\beta$ such that the block, once started by hand, preserves a uniform motion down the plane, then, since $\tan \beta_m = f$ for friction of motion, $\beta_m$ may be less than the $\beta$ in Fig. 168, for friction of rest. 159. Another apparatus consists of a horizontal plane, a pulley, cord, and two weights, as shown in Fig. 59. The masses of the cord and pulley being small and hence neglected, the analysis of the problem when $G$ is so large as to cause an accelerated motion is the same as in that example (2) in § 57), except in Fig. 60, where the frictional resistance $fN$ should be put in pointing toward the left. $N$ still = $G$, and . . . $$S - fG = (G + g)p; \quad \quad \quad \quad \quad \quad \quad (1)$$ while for the other free body in Fig. 61 we have, as before, $$G - S = (G + g)p; \quad \quad \quad \quad \quad \quad \quad (2)$$ From (1) and (2), $S$ the cord-tension can be eliminated, and solving for $p$, writing it equal to $2a + t$, $s$ and $t$ being the observed distance described (from rest) and corresponding time, we have finally for friction of motion $$f = \frac{G}{G_i} - \frac{G + G_i}{G_i} \cdot \frac{2a}{gt}; \quad \quad \quad \quad \quad (3)$$ If $G$, Fig. 59, is made just sufficient to start the block, or sledge, $G_o$, we have for the friction of rest $$f = \frac{G}{G_i}; \quad \quad \quad \quad \quad \quad \quad (4)$$ 160. Results of Experiments on Sliding Friction.--Professor Tissotin in his article on Friction (which the student will do well to read) in Johnson's Cyclopaedia gives the following epitome of results from General Morin's experiments (made for the French Government in 1838): A diagram showing a horizontal plane with a block sliding on it, connected to a pulley system with weights.
MECHANICS OF ENGINEERING.
TABLE FOR FRICTION OF MOTION.
No. Surfaces Uqquent. Angle $\phi$ $f = \tan \phi$
1 Wood on wood. None. 14° to 36° 0.35 to 0.50
2 Wood on metal. None. 36° to 51° 0.50 to 0.60
3 Metal on wood. None. 51° to 80° 0.60 to 0.85
4 Metal on metal. Soap. 11° 0.20
5 Leather on wood. None. 29° to 47° 0.56
6 Metal on leather. Soap. 47° to 67° 0.36
7 Metal on metal. Water. 20° 0.36
8 Leather on metal. Oil. 4° to 12° 0.15
9 Smoother and best lubricated surfaces.
f't' to t"f" to f'f' to f"f" to f'
                                                                    0.85 to 0.086 For friction of rest, about 40% may be added to the coefficients in the above table. In dealing with stone blocks of an arching, $\varphi$ is commonly taken with $30^\circ$, i.e., $f = \tan 30^\circ = 0.58$ as a low safe value; it is considered that if the direction of pressure between two stones makes an angle $>30^\circ$ with the normal to the joint (see §161) slipping may take place (the adhesion of cement being neglected). General Morin states that for a sledge on dry ground $f =$ about $0.66$. Weisbach gives for metal on metal, dry (R.R brakes for example), $f =$ from $0.15$ to $0.94$. Trautwine's Pocket-Book gives values of $f$ for numerous cases of friction. **161. Cone of Friction.** — Fig. 170. Let $A$ and $B$ be two rough blocks, of which $B$ is immovable, and $P$ the resultant of all the forces acting on $A$, except the pressure from $B$. $B$ can furnish any required normal pressure $N$ to balance $P$ cos $\beta$, but the limit of its tangential resistance is $fPN$, by the law of friction, or in other words, so long as the line of action of $P$ is within the "cone of friction" generated by revolving $OC$ about $OA$, the block $A$ will not A diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Caption: Diagram showing a rough block B resting on a smooth block A, with a force P applied at point M. The angle between P and OA is denoted as β, and the angle between P and OB is denoted as φ. The cone of friction is shown as a curved surface around point O, with φ as the apex angle. Description: A rough block B rests on a smooth block A, with **FRICTION.** 160 slip on $B$, and the tangential resistance of $B$ is simply $P \sin \beta$ but if $\beta > \varphi$, this tangential resistance being only $fN$ and $< P \sin \alpha$, $A$ will begin to slip with an acceleration. 168. Problems in Sliding Friction.—In the following problems $f$ is supposed known at points where rubbing occurs, or is impending. As to the pressure $N$ to which the friction is due, it is generally to be considered unknown until determined by the conditions of the problem. Sometimes it may be an advantage to deal with the single unknown force $P$ (resultant of $N$ and $fN$) acting in a line making the known angle $\varphi$ with the normal (on the side away from the motion). Problem 1.—Required the value of the weight $P$, Fig. 171, the slight inclination to which it causes motion of the horizontal rod $OB$, resting through planes at $45^{\circ}$. The weight $G$ of the rod may be applied at the middle. Consider the rod free; at each point of contact there is an unknown $N$ and a friction due to it $fN$; the tension in the cord will be $T = P$, since there is no acceleration and no friction at pulley. Notice that the direction of the friction both opposing the impeding motion. [The student should not rush to the conclusion that $N$ and $N_1$ are equal, and are the same as would be produced by the components of $G$ if the latter were transferred to $A$ and resolved along $AO$ and $AB$; but should await the legitimate results deduced by algebra, from the equations of condition for the equilibrium of a system of forces in a plane. Few problems in Mechanics are so simple as to admit of an immediate mental solution on inspection; and guess-work should be carefully avoided.] Taking an origin and two axes as in figure, we have (equ. (3), § 36), denoting the sine of $45^{\circ}$ by $m$, $$X \ldots fN + mG - N = 0; \quad . . . (1)$$ $$Y \ldots N + fN - mg = 0; \quad . . . (2)$$ $$X(Pa) \ldots fNa + Na - Gb = 0; \quad . . . (3)$$ 170 MECHANICS OF ENGINEERING. The three unknowns $P$, $N$, and $N_1$ can now be found. Divide (3) by $a$, remembering that $b = m$, and solve for $N$; substitute it in (2) and $N_1$ also becomes known; while $P$ is then found from (1) and is $$P = \frac{2mgf}{1+f} = \frac{f\sqrt{g}}{1+f} \cdot G.$$ PROBLEM 2.—Fig. 172. A rod, centre of gravity at middle, leans against a rough wall, and rests on an equally rough floor; how small may the angle $\alpha$ become before it slips? Let $a$ be the half-length. The figure shows the rod free, and following the suggestion of §162, a single unknown force $P$, making a known angle $\varphi$ (whose tan $= f$) with the normal $DE$, is put in at $D$, leaning away from the direction of the impending motion, instead of an $N$ and $fN$; similarly for the other two forces, the most direct method of finding without introducing the other two unknowns $P_1$ and $P_2$ at all, is to use the principle that if three forces balance, their lines of action must intersect in a point. That is, $P$, must intersect the vertical containing $G$, the weight, in the same point as $P_1$, viz., $A$. Now $EA$, and also $EC = a \cos \alpha$, $\therefore ED = a \cos \alpha \cot \varphi$ and $AB = a \cos \alpha \tan \varphi.$ But $DF$, which is $2a \sin \alpha = DE - AB;$ $\therefore 2a \sin \alpha = a \cos \alpha [\cot \varphi - \tan \varphi].$ (1) Dividing by $\cos \alpha$, and noting that $\tan \varphi = f = 1 + \cot \varphi,$ we obtain for the required value of $\alpha$ $$\tan \alpha = \frac{1}{2} \frac{1-f^2}{f};$$ and finally, $\tan \alpha = \cot 2\varphi,$ after some trigonometrical reduction. That is, $\alpha$ is the complement of double the angle of friction. **PROBLEM 3.**—Fig. 173. Given the resistance $Q$, acting parallel to the fixed guide $C$, the angle $\alpha$, and the (equal) coefficients of friction at the rubbing surfaces, required the amount of the horizontal force $P$, at the head of the block $A$ (or wedge), to overcome $Q$ and the frictions. $D$ is fixed, and $ab$ is perpendicular to $cd$. Here we have four unknowns, viz., $P$, and the three pressures $N_1$, $N_2$, and $N_3$ between the blocks. Consider $A$ and $B$ as free bodies, separately (see Fig. 174), remembering Newton's law of action and reaction. The full values (e.g., $f/N$) of the frictions are put in, since we suppose a slow movement. For $A$, $\Sigma X = 0$ and $\Sigma Y = 0$ give $$N_1 - N \cos \alpha + fN \sin \alpha - P \sin \alpha = 0; \quad (1)$$ $$fN_1 + N \sin \alpha + fN \cos \alpha - P \cos \alpha = 0; \quad (2)$$ For $B$, $\Sigma X$ and $\Sigma Y$ give $$Q - N_1 + fN_1 = 0; \quad (3)$$ $$N_1 - fN_1 = 0; \quad (4)$$ Solve (4) for $N_1$ and substitute in (3), whence $$N_1(1-f^2) = Q; \quad (5)$$ Solve (2) for $N$, substitute the result in (1), as also the value of $N_1$ from (5), and the resulting equation contains but one unknown, $P$. Solving for $P$, putting for brevity $$f\cos \alpha + sin \alpha = m \quad and \quad cos \alpha - f\sin \alpha = n,$$ we have $$P = \frac{(m+n)fQ}{(n,\cos\alpha+m,\sin\alpha)(1-f^2)}; \quad (6)$$ or $$P = Q(m+fn)/(1-f^2). \quad (7)$$ Fig. 173. Fig. 174. 171 172 MECHANICS OF ENGINEERING. Numerical Example of Problem 3.—If $Q = 120$ lbs., $f = 0.80$ (an abstract number, and : the same in any system of units), while $\alpha = 14^{\circ}$, whose sine is $0.240$ and cosine is $.970$, then $$m = 0.2 \times .97 + .34 = .43 \quad n = .97 - .2 \times .34 = .92,$$ whence $$P = 0.64 f = 76.8 \text{ lbs}.$$ While the wedge moves 2 inches $P$ does the work (or exerts an energy) of $2 \times 76.80 = 153.6$ in.-lbs. = 12.8 ft.-lbs. For a distance of 2 inches described by the wedge horizontally, the block $B$ (and : the resistance $Q$) has been moved through a distance $= 2 \times sin 14^{\circ} = 0.48$ in. along the guide $C$, and hence the work of $120 \times .048 = 57.6$ in.-lbs. has been done upon $Q$. Therefore for the supposed portion of the motion $153.6 - 57.6 = 96.0$ in.-lbs. of work has been lost in friction (converted into heat). It is noticeable that if $\alpha$ should be $90^{\circ}$, $P = Q$, and there is no friction (the weights of the blocks have been neglected). Problem 4.—Numerical.—With what minimum pressure $P$ should the pulley $A$ be held against $B$, which it drives by "frictional gearing," to transmit 9 H.P., if $\alpha = 45^{\circ}$, for impending relative motion, i.e., for impending slipping = 0.40, and the velocity of the pulley-rim is 9 ft. per second? The limit-value of the tangential "grip" $$T = 9H.P. \times .40 \times P \sin 45^{\circ},$$ $$9 H.P. = 2 \times .550 = 1100 \text{ ft}-\text{lbs}. \text{ per second}.$$ Putting $T \times X$ ft. = 1100, we have $$2 \times .40 \times \sqrt{2} \times P \times X = 1100 ; \therefore P = .215 \text{ lbs}.$$ Problem 5.—A block of weight $G$ lies on a rough plane, inclined an angle $\beta$ from the horizontal; find the pull $P$, making an angle $\alpha$ with the first plane, which will maintain a uniform motion up the plane. A diagram showing a block on a inclined plane with a pulley attached to it. Fig. No. 5 **FRICITION.** 173 **PROBLEM 7—Same as 6, except that the pull $P$ is to permit a uniform motion down the plane.** **PROBLEM 8.—The thrust of a screw-prowel is 15 tons. The radius of the screw-prowel exerted has a mean radius of 8 inches, the shaft makes one revolution per second, and $f = 0.06$. Required the H. P. lost in friction from this cause.** Ans. 13.7 H. P. 163. The Bent-Lever with Friction. Worn Bearing.—Fig. 176. Neglect the weight of the lever, and suppose the plumber-block so worn that there is contact between one element only of the shafts. Given the amount and line of action of the resistance $R$, and the line of action of $P$, required the amount of latter for impending slipping in the direction of the dotted arrow. As $P$ gradually increases, the shaft of the lever (or gear-wheel) rolls on its bearing, till the line of contact has reached some position $A$, when rolling ceases, and the lever begins to find $P$ equal to the value of $P$, note that the total action of the bending upon the lever is some force $P_1$, applied at $A$ and making a known angle $\varphi (\varphi = \tan \theta)$ with the normal $AC$. $P_1$ must be equal and opposite to the resultant of the known $R$ and the unknown $P$, and hence graphically (a graphic is much simpler here than an analytical solution) if we describe about $OA$ circle of radius $r = \sin \varphi$, $r$ being the radius of shaft (or wedge), and draw a tangent at $D$, then by drawing a line parallel to $AD$ at $DG$ made $= R$, to scale, and $GF$ drawn parallel to $D\ldots P$, $P$ is determined, being $DE$, while $P_1 = DEF$. If the known force $R$ is capable of acting as a working force, by drawing the other tangent $DB$ from $D$ to the "friction-circle," we have $P = DH$, and $P_1 = DK$, for impending rotation in an opposite direction. If $R$ and $P$ are the tooth-pressure upon two spur-wheels, keyed upon the same shaft and nearly in the same plane, the 174 MECHANICS OF ENGINEERING. same constructions look good, and for a continuous uniform motion, since the friction $=P_r \sin \varphi$, the work lost in friction is $\frac{1}{2} = [P_r \sin \varphi] 2\pi r$ per revolution. It is to be remarked that without friction $P_r$ would pass through $C$, and that the moments of $R$ and $P_r$ would balance about $O$ (for rest or uniform rotation); whereas with friction they balance about the proper tangent-point of the friction-circle. Another way of stating this is as follows: So long as the resultant of $P$ and $R$ falls within the "dead-angle" $BDA$, motion is impossible in either direction. If the weight of the lever is considered, the resultant of it and the force $R$ can be substituted for the latter in the foregoing. 164. Bent-Lever with Friction. Triangular Bearing.—Like the preceding, the gudgeon is much exaggerated in the figure (177). For impending rotation in direction of the force $P_r$, the total moment about $A_1$, summing in known directions, making angles $=\varphi$ with the respective normals, and inclined away from the slipping. Join the intersections $D$ and $L$. Since the resultant of $P$ and $R$ at $D$ must act along $DL$ to balance that of $P_r$ and $P_s$, having given one force, say $P_r$, we easily find $F = DE$, while $P_s = LM$ and $LN$ respectively; $LO$ having been made equal to $DF$, and the parallelogram completed. (If the direction of impending rotation is reversed, the change in the construction is obvious.) If $P_r = 0$, the case reduces to that in Fig. 176; if the construction gives $P_r$ negative, the supposed contact at $A_1$ is not realized, and the angle $\angle A_1CA_2$ should be increased, or shifted, until $\angle P_1$ is positive. As before, $P$ and $R$ may be the tooth-pressures on two A diagram showing a triangular bearing with forces applied at points D, E, F, G, H, I, J, K, L. Fig. 177 **FRICTION.** 175 spur-wheels nearly in the same plane and on the same shaft; if so, then, for a uniform rotation. Work lost in fric. per revol. = $P_1 \sin \phi + P_2 \sin \phi$ per r. 165. **Axle-Fricion.** —The two foregoing articles are intro- ductory to the subject of axle-friction. When the bearing is new, or nearly so, the elements of the axe which are in contact with the bearing are infinite in number, thus giving an infinite number of unknown forces similar to $P_1$ and $P_2$ of the last paragraph. The laws of distribution of these forces and similar theories as to the law of distribution of these pressures are of little use, considering the uncertainties as to the value of $f(\tan \phi)$; hence for practical purposes axle-friction may be written $$F = fR,$$ in which $f$ is a coefficient of axle-friction derivable from experiments with axes, and $R$ the resultant pressure on the bearing. In some cases $R$ may be partly due to the tightness of the belt with which it is driven, this being neglected. As before, the work lost in overcoming axle-friction per revolution = $fR2\pi$, in which $r$ is the radius of the axe. $f$, like $f$, is an abstract number. As in Fig. 176, a "friction-circle," of radius $=fr$, may be considered as subtending the "dead-angle." 166. **Experiments with Axle-Fricion.** —Prominent among recent experiments have been those of Professor Thurston (1872-73), who invented a special instrument for that purpose, shown (in principle only) in Fig. 178. By means of an internal spring, the amount of whose compression is read off by a weighted bar or pendulum is caused to exert pressure on a projecting axe from which it is suspended. The axe is made to rotate at any desired velocity by some source of power, the axle-friction causing A diagram showing a friction circle and a weight attached to a pendulum. Fig. 178 176 MECHANICS OF ENGINEERING. the pendulum to remain at rest at some angle of deviation from the vertical. The figure shows the pendulum for the action of gravity upon it being $G$, that of the axle consisting of the two pressures, each $R$, and of the two frictions (each being $F = f R$), due to them. Taking moments about $C$, we have for equilibrium $$9f^2 Rr = Gh,$$ in which all the quantities except $f$ are known or observed. The temperature of the bearing is also noted, with reference to its different lubricants employed. Thus the instrument covers a wide range of relations. General Morin's experiments as interpreted by Weibach give the following practical results: For iron axles, in iron or brass bearings $$f' = \left\{ \begin{array}{ll} 0.054 & \text{for well-sustained lubrication;} \\ 0.07 & \text{to 0.08 for ordinary lubrication.} \end{array} \right.$$ By "pressure per square inch on the bearing" is commonly meant the quotient of the total pressure in lbs. by the area in square inches obtained by multiplying the width of the axle by the length of bearing (this length is quite commonly four times the diameter); call it $p$, and the velocity of rubbing in feet per minute, $v$. Then, according to Rankine, to prevent overheating, we should have $$p(v + 20) < 4480 \ldots \quad (\text{not homog.)}$$ Still, in marine-engine bearings $pv$ alone often reaches 60,000, as also in some locomotives (Cutlerill). Good practice keeps $p$ within the limit of 800 (lbs. per sq.-in.) for other metals than steel (Thurston), for which 1300 is sometimes allowed. With $v = 200$ (feet per min.) Professor Thurston found that for ordinary lubricants $p$ should not exceed values ranging from 30 to 75 (lbs. per sq. in.). The product $pv$ is obviously proportional to the power expended in wearing the rubbing surfaces, per unit of area. **FRICITION.** 177 **167. Friction-Wheels.** — A single example of their use will be given, with some approximations to avoid complexity. Fig. 179. $G$ is the weight of a heavy wheel, $P$ is a known vertical resistance (tooth-pressure), and $P'$ is an unknown vertical working force, whose value is to be determined to maintain a uniform rotation. The utility of the friction-wheels is also to be shown. The resultant of $F_{a}$, $G$, and $P'$ is a vector $R$, acting nearly through the centre $C$ of the main axle which rolls on the four friction-wheels. $R$, resolved along $CA$ and $CB$, produces (nearly) equal pressures, each being $N = R + 2 \cos \alpha$, at the two axles of the friction-wheels, which rub against their fixed plumber-blocks. $R = P + P' + G$, and $...$ contains the unknown $P'$, but approximately $P' = N - P - G$, i.e., is nearly the same (in this case) when friction-wheels are employed. When $G$ makes one revolution, the friction $f'N$ at each axle $C_i$ is overcome through a distance of $(r_i : a_i) 2\pi r$, and Work lost per revol. with friction-wheels, $$= f'N \frac{r_i}{a_i} 2\pi r = \frac{r_i}{a_i} \cos \alpha f' R 2\pi r.$$ Whereas, if $C$ revolved in a fixed bearing, Work lost per revol. without friction-wheels, $$= f' R 2\pi r.$$ Apparently, then, there is a saving of work in the ratio $r_i : a_i$, cos $\alpha$, but strictly the $R$ is not quite the same in the two cases ; for with friction-wheels the force $P'$ is less than without, and $R$ depends on $P$ as well as on the known $G$ and $P'$. By diminishing the angle $\alpha$, $r_i : a_i$, and the angle $\alpha$, the saving is increased. If $\alpha$ were so large that cos $\alpha < r_i : a_i$, there would be no saving, but the reverse. As to the value of $P'$ to maintain uniform rotation, we have Fig. 179. 178 MECHANICS OF ENGINEERING. for equilibrium of moments about $C$, with friction-wheels (considering the large wheel and axle free), $$Pb = P\delta + 2Tr \quad \ldots \quad (1)$$ in which $T$ is the tangential action, or "grip" between one pair of friction-wheel and the axis $C$ which rolls upon them. $T$ would not equal $fN$ until slipping took place or were impending at $E$, but is known by considering a pair of friction-wheels free, when $\Sigma(Pa)$ about $C$, gives $$Ta = fNr = f\frac{R}{2}\cos a'$$ which in (1) gives finally $$P = \frac{b}{2}P + \frac{r}{a}\cos a'f\frac{R}{b} \quad \ldots \quad (2)$$ Without friction-wheels, we would have $$P = \frac{b}{2}P + f\frac{R}{b} \quad \ldots \quad (3)$$ The last term in (2) is seen to be less than that in (3) (unless $\alpha$ is too large), in the same ratio as already found for the saving of work, supposing the $fRa'$ equal. If $P$, were on the same side of $C$ as $P$, it would be of an opposite direction, and the pressure $R$ would be diminished. Again, if $P$ were horizontal, $R$ would not be vertical, and the friction-wheel axes would not bear equal pressures. Since $P$ depends on $P$, $G$, and the frictions, while the friction depends on $R$, and $R$ on $P$, $G$, and $P$, an exact analysis is quite complex and is not warranted by its practical utility. Example.--A car weighing vertically 25,000 lbs., requiring the force $P$ to be applied at its circumference to maintain a uniform motion, with $a = 15$ ft., and $r = 5$ inches. Here $P_1 = 0$, and $R = G$ (nearly); neglecting the influence of $P$ on $R$), i.e., $R = 25,000$ lbs. First, without friction-wheels (adopting the foot-pound-second system of units), with $f = 0.7$ (abstract number). From eq. (3) we have $$P = 0 + 0.07 \times 25,000 \times (\frac{1}{15} - 1) = 48.6 \text{ lbs.}$$ **FRICTION.** 179 The work lost in friction per revolution is $$fR\varpi = 0.07 \times 25,000 \times 3.14 \times \frac{1}{2} = 4580 \text{ ft-lbs.}$$ Secondly, with friction-wheels, in which $r : a = \frac{1}{2}$ and $\cos \alpha = 0.80$ (i.e., $\alpha = 36^\circ$). From eq. (2) $$P = 0 + \frac{1}{2} \times 45.6 = 22.8 \text{ lbs.},$$ while the work lost per revolution $$= \frac{1}{2} \times 4580 = 1145 \text{ ft-lbs.}$$ Of course with friction-wheels the wheel is not so steady as without. In this example the force $P$ has been simply enough to overcome friction. In case the wheel is in actual use, $P$ is the weight of water actually in the buckets at any instant, and does the work of overcoming $P_r$, the resistance of the mill machinery, and also the friction. By placing $P$, pointing upward on the same side of $C$ as $P_r$, and making $b_r$ nearly $b$, $R$ will be $G$ nearly, just as when the wheel is running empty; and the foregoing numerical results will still hold good for practical purposes. **168. Friction of Pivots.—In the case of a vertical shaft or axle, and sometimes in other cases, the extremity requires support against a thrust along the axis of the axe or pivot. If the end of the pivot is flat and also the surface against which it rubs, we may consider the pressure, and therefore the friction, as uniform over the surface. With a flat circular pivot, then, if $R$ be the resultant of all forces whose resultant is equal to their sum, and is applied a distance of $r$ from the centre. Hence the sum of the moments of all the frictions about the centre $= fR\varpi$, in which $R$ is the axial pressure. Therefore a force $P$ necessary to overcome the friction with uniform rotation must have a moment $$Pa = fR\varpi.$$ 180 MECHANICS OF ENGINEERING. and the work lost in friction per revolution is $$fRr\cdot\frac{d}{dt} = \frac{1}{2}frRr \quad \ldots \quad (1)$$ As the pivot and step become worn, the resultant friction in the small sector probably approach the centre; for the greatest wear occurs first near the outer edge, since there the product $pr$ is greatest (see § 16). Hence for $\frac{d}{dt}$ we may more reasonably put $\frac{d}{dt}$. Example.—A vertical flat-ended pivot presses its step with a force of 12 tons, is 6 inches in diameter, and makes 40 revolutions per minute. Required the H. P. absorbed by the friction. Supposing the pivot and step new, and $f$ for good lubrication = 0.07, we have, from eq. (1) ($\frac{1}{2}$foot-lb.second), Work lost per revolution $$= .07 \times 24,000 \times 6.38 \times \frac{1}{2} = 1758.4 \text{ ft.-lbs.,}$$ and . . . work per second $$= 1758.4 \times \frac{1}{2} = 1172.9 \text{ ft.-lbs.,}$$ which + 550 gives 2,13 I. P. absorbed in friction. If ordinary axle-friction also occurs its effect must be added. If the flat-ended pivot is hollow, with radii $r_1$ and $r_2$, we may put $\frac{d}{dt}(r_1 + r_2)$ instead of the $\frac{d}{dt}$ of the preceding. It is obvious that the smaller the lever-arm given to the resultant friction in each sector of the rubbing surface the smaller the power lost in friction. Hence pivots should be made as large as possible, with strength. For a conical pivot and step, Fig. 18, the resultant friction in each sector of the conical bearing surface has a lever-arm = $\frac{d}{dt}$ about the axis $A$, and a value > than for a flat-ended pivot; for, on account of the wedge-like action of the bodies, the pressure causing friction is greater. The sum of the moments of these resultant frictions about $A$ is the same as if only two elements of the cone received pressure (each $N = \frac{1}{2}B + sin\alpha$). Hence the **FRICTION.** 181 moment of friction of the pivot, i.e., the moment of the force necessary to maintain uniform rotation is $$Pa = f^2 N \frac{R}{2} = f_1 \frac{R}{2} \frac{9}{\sin \alpha}$$ and work lost per revolution $$= \frac{4}{3} \sqrt{\frac{R}{\sin \alpha}}$$. By making $r$, small enough, these values may be made less than those for a flat-ended pivot of the same diameter = 2$r$. In Schiele's "anti-friction" pivot the outline is designed according to the following theory for securing uniform vertical wear. Let $p$ = the pressure per horizontal unit of area (i.e., $= R +$ horizontal projection of the actual rubbing surface); this is assumed constant. Let the unit of area, $A$, equal, for algebraic simplicity. The friction on the rubbing surface, whose horizontal projection = unity, is $= fN = f(p + \sin \alpha)$ (see Fig. 183; the horizontal component of $p$ is annulled by a corresponding one opposite). The work per revolution in producing wear on this area is $fN2\pi r$. But the vertical depth of wear per revolution is to be the same at all parts of the surface; and this implies that the same volume of material is worn away under each horizontal unit of area. Hence $fN2\pi$, i.e., $\int_{0}^{2\pi} fN2\pi r$, is to be constant for all values of $y$; and since $fp$ and $2\pi$ are constant, we must have, as the law of the curve, $$\frac{y}{\sin \alpha}$$ I.e., the tangent $BC =$ the same at all points. This curve is called the "tractrix." Schiele's pivots give a very uniform wear at high speeds. The smoothness of wear prevents leakage in the case of cocks and fancoats. **186. Normal Pressure of Belting.—When a perfectly flexible cord, or belt, is stretched over a smooth cylinder, both at rest, A diagram showing a belt tensioned around a cylinder. The belt is shown with a slight upward curvature due to tension. The normal pressure $p$ is indicated along the belt, and its horizontal component $p_A$ is shown separately. 182 MECHANICS OF ENGINEERING. the action between them is normal at every point. As to its amount, $p$, per linear unit of arc, the following will determine. Consider a semi-circle of the cord free, neglecting its weight. Fig. 183. The forces holding it in equilibrium are the tensions at the two ends (these are equal); manifestly, the cylinder being symmetrical, for they are the only two forces having more than about $\frac{1}{2}C$ each (the same lever-arm), and the normal pressures, which are infinite in number, but have an intensity $p$, per linear unit, which must be constant along the curve since $S$ is the same at all points. The normal pressure on a single element, $ds$, of the cord is $=pds$, and its $X$ component is $=pds\cos\theta=pds\sin\theta$. Hence $2X=0$ gives $$rp_{\frac{d\theta}{dr}}\cos\theta d\theta-2S=0,\mathrm{i.e.,}\;rp_{\frac{d\theta}{dr}}\sin\theta=2S;$$ $$:rp(1-(-1))=2S \mathrm{~or~} p=\frac{S}{r}. \quad \ldots \quad (1)$$ 170. Belt on Rough Cylinder. Impending Slipping.—If friction is possible between the two bodies, the tension may vary along the arc of contact, so that $p$ also varies, and consequently Fig. 184. Fig. 185. the friction on an element $ds$ being $=f(pds)=f(S+r)ds$, also varies. If slipping is impending, the law of variation of the tension $S$ may be found, as follows : Fig. 184, in which the FRICTION. 183 impending slipping is toward the left, shows the cord free. For any element $da$, of the cord, we have, putting $\mathscr{B}$ (moment about $O$) = 0 (Fig. 185), \[ (S + dS)d\theta = Sdr + dFr; \text{i.e., } dF = dS, \] or (see above) $dS = f(S + pdS)$. But $ds = rd\theta$; hence, after transforming, $$f\frac{dS}{S} = \frac{dS}{S} \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1) $$ In (1) the two variables $\theta$ and $S$ are separated; (1) is therefore ready for integration. $$\therefore f\int_{a}^{b} d\theta = J_{a,b}\frac{dS}{S}; i.e.,$$ $$f\alpha = log.S_{a} - log.S_{b} = log.\left[\frac{S_{b}}{S_{a}}\right]$$ (9) Or, by inversion, $$S_{a}e^{\alpha} = S_{b}$$ (3) $\epsilon$, denoting the Napierian base, $= 2.71828 +$, $\alpha$ of course is in $\pi$-measure. Since $S_a$ evidently increases very rapidly as $\alpha$ becomes larger, $S_b$ remaining the same, we have the explanation of the well-known fact that a comparatively small tension, $S_a$, exerted by a man, is able to prevent the slipping of a rope around a pulley-beam when the further end is under the great tension $S_b$ due to the stopping of a moving steamer. For example, with $f = 1$, we have (Weisbach) for $\alpha = 1$ turn, or $\alpha = 2\pi$, $S_a = 16.98X$; for $\alpha = 2$ turns, or $\alpha = 4\pi$, $S_a = 2.858X$; for $\alpha = 3$ turns, or $\alpha = 6\pi$, $S_a = 8.192X$; for $\alpha = 4$ turns, or $\alpha = 8\pi$, $S_a = 65.94X$; for $\alpha = 5$ turns, or $\alpha = 10\pi$, $S_a = 4348.56X$. If slipping actually occurs, we must use a value of $f$ for friction of motion. Example.--A leather belt drives an iron pulley, covering one half the circumference. What is the limiting value of the 184 MECHANICS OF ENGINEERING. ratio of $S_1$ (tension on driving-side) to $S_2$ (tension on following side) if the belt is not to slip, taking the low value of $f = 0.25$ for leather on iron ! We have given $f \times \pi = 0.25 \times \pi = 0.7854$, which by eq. (2) is the Napierian log. of $(S_1 : S_2)$ when slipping occurs. Hence the common log. of $(S_1 : S_2) = 0.7854 \times 0.43429 = 0.34109$ ; i.e., if $$\log(S_1 : S_2) = 2.193,$$ say 2.2, the belt will slip (for $f = 0.25$). (0.43429 is the modulus of the common system of logarithms, and $\pi = 2.30385$. See example in § 45.) As in § 169, the relation between $S_1$ and $S_2$ is not strictly true, since the tensions at the two ends of an element $ds$ are partly employed in furnishing the necessary deviating force to keep the element of the cord in its circular path, the remainder producing normal pressure. 171. Transmission of Power by Belting or Wire Rope.—In the simple design in Fig. 186, it is required to find the motive weight $G$, necessary to overcome the given resistance $R$ at a uniform velocity $v$, also the proper stationary tension weight $G_s$ to prevent slipping of the belt on its pulleys, and the amount of power, $L$, transmitted. In other words, Given : $$\begin{cases} R, a, r_a, r_s ; \alpha = \pi \text{ for both pulleys;} \\ v; \text{and } f \text{ for both pulleys;} \end{cases}$$ Required : $$\begin{cases} G_s: G_s \text{ to furnish } L: G_s \text{ for no slip;} \\ v: \text{the velocity of } G: v' \text{ that of belt;} \\ \text{and the tensions in belt}. \end{cases}$$ Fig. 186. **FRICTION.** 185 Neglecting axle-friction and the rigidity of the belting, the power transmitted is that required to overcome $R$ through a distance = $v$, every second, i.e., $$L = R v_n \dots \dots \dots (1)$$ Since (if the belts do not slip) $$a : r :: e' : v, \text{ and } a_r : r :: e' : v_n,$$ we have $$e' = \frac{a}{r} v_n, \text{ and } v = \frac{r}{a} e' v_n \dots \dots (2)$$ Neglecting the mass of the belt, and assuming that each pulley revolves on a gravity-axe, we obtain the following, by considering the free bodies in Fig. 187: A diagram showing two pulleys connected by a belt. The pulleys are labeled S_1 and S_2. The belt is shown in two positions: one where it is free to move (a free), and another where it is tuck (b and tuck free). The pulleys are labeled G_1 and G_2. (6 free) (6 and tuck free) Fig. 187. $$\mathbb{F}(\text{moms.}) = 0 \text{ in } A_f \text{free gives } Rr = (S_n - S_s)a ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ $$\mathbb{F}(\text{moms.}) = 0 \text{ in } B \text{free gives } Gr = (S_n - S_s)a ; . . . (4)$$ whence we readily find $$G = \frac{a}{r} \frac{r_s}{a_s} R.$$ Evidently $R$ and $G$ are inversely proportional to their velocities $v_s$ and $v_t$; see (2). This ought to be true since in Fig. 186 $G$ is the only working-force, $R$ the only resistance, and the motions are uniform; hence (from eq. (XVI), § 148) $$Gv - Rv_t = 0.$$ $$\mathbb{F}X = 0, \text{ for } B \text{and tuck free}, \text{ gives}$$ $$G_t = S_s + S_e \dots \dots \dots (5)$$ while, for impending slip, $$S_e = S_sp_n \dots \dots \dots (6)$$ 186 MECHANICS OF ENGINEERING. By elimination between (4), (5), and (6), we obtain $$G_s = G_a \cdot \frac{e^{\alpha} + 1}{e^{\alpha} - 1} = \frac{L}{v} \cdot \frac{e^{\alpha} + 1}{e^{\alpha} - 1} \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \\ (7)$$ and $$S_a = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot e^{-\alpha} = L/v\cdot e^{\alpha}\cdot FRICTION. 187 way to the uniform motion of a roller or wheel rolling upon it corroborate approximately the following theory. The word friction is hardly appropriate in this connection (except when the roadway is perfectly elastic, as will be seen), but is sanctioned by usage. First, let the roadway or track be compressible, but inelastic, $G$ the weight of the roller and its load, and $P$ the horizontal force necessary to produce a uniform motion (both of translation and rotation). The track (or roller itself) being compressed just in front, and not reacting symmetrically from behind, its resultant pressure against the roller is not at $O$ vertically under the centre, but some small distance, $OD = b$, in front. (The successive crushing of small projecting particles has the same effect.) Since for this case of motion the forces have the same relations as if balanced (see § 124), we may put $\mathbf{F}$ mome. about $D = 0$, $$\therefore Pr = Gb; \text{ or, } P = \frac{b}{r}G. \quad \ldots \ldots (1)$$ Coulomb found for Rollers of lignum-vitae on an oak track, $$b = 0.0189 \text{ inches};$$ Rollers of elm on an oak track, $$b = 0.0320 \text{ inches}.$$ Weibach's experiments give, for cast-iron wheels 20 inches in diameter on cast-iron rails, $$b = 0.0183 \text{ inches};$$ and Bittinger, for the same, $$b = 0.0193 \text{ inches}.$$ Pambour gives, for iron railroad wheels 39.4 inches in diameter, $$b = 0.0196 \text{ to}$$ $$\ldots \ldots 0.0216 \text{ inches}.$$ According to the foregoing theory, $P$, the "rolling friction" (see eq. (1)), is directly proportional to $G$, and inversely to the radius, if $b$ is constant. The experiments of General Morin and others confirm this law while those of Dupont, Poiseuille, and Sauvage indicate it to be proportional directly to $G$, and inversely to the square root of the radius. Fig. 198. CLAR 188 MECHANICS OF ENGINEERING. Although $b$ is a distance to be expressed in linear units, and not an abstract number like the $f$ and $f'$ for sliding and axial friction, it is sometimes called a "coefficient of rolling friction." In eq. (1), $b$ and $r$ should be expressed in the same unit. Of course if $P$ is applied at the top of the roller its lever-arm about $D$ is $\vartheta r$ instead of $r$, with a corresponding change in eq. (1). With ordinary railroad cars the resistance due to axle and rolling frictions combined is about 8 lbs. per ton of weight on a level track. For wagons on macadamized roads $b = \frac{3}{4}$ inch, but on soft ground from 2 to 3 inches. Secondly, when the roadway is perfectly elastic. This is chiefly of theoretic interest, since at first sight no force would be considered necessary to maintain a uniform rolling motion. But, as the material of the roadway is compressed under the roller its surface is first elongated and then recovers its former state; hence some rubbing and consequent sliding friction must occur. Fig. 189 gives an exaggerated view of the circumstances, $P$ being the horizontal force applied at the centre necessary to maintain a uniform motion. The roadway (rubber for instance) is heaped up both in front and behind the roller, $O$ being the point of greatest pressure and elongation of the surface. The forces acting are $G$, $P$, the normal pressures, and the frictions due to them, and must form a balanced system. Hence, since $G$ and $P$ act along the normal pressure lines through $C$, the resultant of the frictions must also pass through $C$; therefore the frictions, or tangential actions, on the roller must be some forward and some backward Fig. 189. $$\alpha$$ FRICITION. 189 (and not all in one direction, as seems to be asserted on p. 260 of Costerill's Applied Mechanics, where Professor Reynolds' explanation is cited). The resultant action of the roadway upon the roller acts, then, through some point $D$, a distance $OD = b$ ahead of $O$, and in the direction $DC$, and we have as before, with $D$ as a centre of moments, $$Pr = Gb, \quad \text{or} \quad P = \frac{b}{r}G.$$ If rolling friction is encountered above as well as below the rollers, Fig. 190, the student may easily prove, by considering three separate free bodies, that for uniform motion $$P = \frac{b + b'}{2r}G, \quad \ldots \ldots (9)$$ where $b$ and $b'$ are the respective "coefficients of rolling friction" for the upper and lower contacts. **Example 1.**—If it is found that a train of cars will move uniformly down a slope due to gravity being the only working force, and friction (both rolling and axle), the only resistance, required the coefficient $f'$ of axle-friction, the diameter of all the wheels being $2r = 30$ inches, that of the journals $2a = 3$ inches, taking $b = 0.09$ inch for the rolling friction. Let us use equation (X.V.) ($\S$ 142), noting that while the train moves a distance $x$ measured on the incline, its weight $G$ does the work $\frac{1}{900}Gx$, the rolling friction $\frac{a}{r}G$ (at the axles) has been overcome through the distance $e$, and the axle-friction (total) through the (relative) distance $\frac{a}{r}$ in the journal boxes; whence, the change in kinetic energy being zero, $$\frac{1}{900}Ga - \frac{a}{r}Ge - \frac{a}{r}Ge = 0.$$ $Ge$ cancels out, the ratios $b : e$ and $a : r$ are $= x : y_1$ and $y_1$, respectively (being ratios or abstract numbers they have the) *That is, the ideal resistance at centre of axles and $\parallel$ to the incline, equal to actual rolling resistance.* Fig. 190. 190 MECHANICS OF ENGINEERING. same numerical values, whether the inch or foot is used), and solving, we have $$f' = 0.05 - 0.0133 = 0.036.$$ Example 2.—How many pounds of tractive effort per ton of load would the train in Example 1 require for uniform motion on a level track? Ans. 10 lbs. 173. Railroad Brakes.—During the uniform motion of a railroad car the tangential action between the track and each wheel is small. Thus, in Example 1, just cited, if ten cars of eight tons each move at the rate of one mile per hour, the backward tangential action of the rails upon each wheel is only 25 lbs. When the brakes are applied to stop the train this action is much increased, and is the only agency by which the rails can retard the train, directly or indirectly; directly, when the pressure of the brakes is so great as to prevent the wheels from turning, thereby causing them to "skid" (i.e., slide) on the rails; indirectly, when the brake-pressure is of such a value as still to permit perfect rolling of the wheel, in which case the rubbing (and heating) between the brake and wheel, and the tendency of the wheel to turn has a value equal or less than the friction of rest. In the first case, then (skidding), the retarding influence of the rails is the friction of motion between rail and wheel; in the second, a force which may be made as great as the friction of rest between rail and wheel. Hence, aside from the fact that skidding produces objectionable flat places on the wheel-tread, the brakes are more effective if so applied that skidding is impending, but not actually produced; for the retarding effect is nearly greater than that of actual slidding (§ 160). This has been proved experimentally by English tests; but retarding effect of axle and rolling friction has been neglected in the above theory. Example 1.—A twenty-ton car with an initial velocity of 80 feet per second (nearly a mile a minute) is to be stopped on a level within 1000 feet; required the necessary friction on each of the eight wheels. Supposing the wheels not to skid, the friction will occur **FRICITION.** 191 between the brakes and wheels, and is overcome through the (relative) distance 1000 feet. Eq. (XVL), § 142, gives (foot- lb.-second system) $$0 - 8F \times 1000 = 0 - \frac{1}{3} \cdot 32.5 \cdot (80)^2,$$ from which $F$ (= friction at circumference of each wheel) = 496 lb. **Example 2.** Suppose skidding to be impending in the fore- going, and the coefficient of friction of rest (i.e., impending slipping) between rail and wheel to be $f = 0.20$. In what distance will the car be stopped? **Ans.** 496 ft. **Example 3.** Suppose the car in Example 1 to be on an up- grade of 60 feet to the mile. (In applying eq. (XVL) here, the weight 20 tons enter as a resistance.) **Ans.** 439 lb. **Example 4.** Consider all four resistances, viz., gravity, rolling friction, and that due to axle friction, the distance being 1000 ft., and $F'$ the unknown quantity. (Take the wheel dimensions of p. 189.) **Ans.** 414 lb. 174. Estimation of Engine and Machinery Friction.—Accord- ing to Professor Cottrell, a convenient way of estimating the work lost in friction in a steam-engine and machinery driven by it is the following : Let $p_m = \text{mean effective steam-pressure per unit of area of piston}$ and conceive this composed of three por- tions, viz, $p_s = \text{the necessary pressure to drive the engine alone un- loaded, at the proper speed}$ $p'_m = \text{pressure necessary to overcome the resistance caused by the useful work of the machines}$; $ep'_m = \text{pressure necessary to overcome the friction of the machinery, and that of the engine over and above its friction when unloaded. This is about } 15\%$ of $p'_m$ (i.e., $e = 0.15$), except in large engines, and then rather less. That is, by formula, $F$ being the piston-area and $l$ the length of stroke, the work per stroke is thus distributed: $$Fp'_m l = F(1 + ep'_m + p'_s);$$ *Recent experiments (1888) by Prof. Thurston show that this surplus engine-friction is practically zero.* 102 MECHANICS OF ENGINEERING. $p$, is "from 1 to 14, or in marine engines 2 lbs. or more per square inch." 175. Anomalies in Friction.—Experiment has shown considerable deviation under certain circumstances from the laws of friction, as stated in § 157 for sliding friction. At pressures below 1 lb. per sq. inch the coefficient $f$ increases when the pressure decreases, while above 300 lbs. (Remm, with iron and steel) it increases with the pressure. With high velocities, however, above 10 ft. per second, $f$ is much smaller than the velocity increases. As for axle-friction, experiments instituted by the Society of Mechanical Engineers in England (see the London Engineer for March 7 and 21, 1884) gave values for $f$ less than $\frac{1}{4}$ when a "bath" of the lubricant was employed. These values diminished with increase of pressure, and increased with the velocity (see below, Hirn's statement). Professor Cotterill says, "It cannot be doubted that for values of $pv$ (see § 160) > 5000 the coefficient of friction of well-lubricated bearings of general construction diminishes with the pressure, but may remain much less than that at low speeds as determined by Morin" (p. 259 of his Applied Mechanics). Professor Thruston's experiments confirmed those of Hirn as to the following relation: "The friction of lubricated surfaces is nearly proportional to the square root of the area and pressure." Hirn also maintained that, "in ordinary machinery, friction varies as the square root of the velocity." 176. Rigidity of Rope.—If a rope or wire cable passes over a pulley or sheave, a force is required on one side greater than that on the other, which must move aside from axle-friction. Since in a given time both $P$ and $Q$ describe the same space, if $P$ is > $Q$, then $P's > Qs$, i.e., the work done by $P$ is > than that expended upon $Q$. This is because some of the work $Ps$ has been expended in bending the stiff rope or cable, and in overcoming friction between the strands, both where the rope passes upon and where it leaves A diagram showing a rope passing over a pulley or sheave. **FRICTION.** 193 the pulley. With hemp ropes, Fig. 191, the material being nearly inelastic, the energy spent in bending it on at $D$ is nearly all lost, and energy must also be spent in straightening. A diagram showing a pulley system with a rope passing over two pulleys. The rope is shown in three different positions: one where it is bent around the first pulley (P), another where it is straightened out (Q), and a third position where it is bent around the second pulley (E). The rope is labeled with letters indicating its position. Fig. 191. It at $E$; but with a wire rope or cable some of this energy is restored by the elasticity of the material. The energy spent in friction or rubbing of strands, however, is lost in both cases. The figure shows geometrically why $P$ must be > $Q$ for a uniform motion, for the lever-arm, $a$, of $P$ is evidently < $b$ that of $Q$. If axle-friction is also considered, we must have $$Pa = Qb + f'(P + Q)r,$$ $r$ being the radius of the journal. Experiments with cordage have been made by Prony, Coulomb, Eyethein, and Weisbach, with considerable variation in the results and formulae proposed. (See Coxe's translation of vol. i., Weisbach's Mechanics.) With pulleys of large diameter the effect of rigidity is very slight. For instance, Weisbach gives an example of a pulley five feet in diameter, on which when the load was 18 lbs. = 1919. A wire rope 1 in. in diameter was used. Of this difference, 19 lbs., only 5 lbs. was due to rigidity, the remainder, 14 lbs., being caused by axle-friction. When a hemp-rope 1.6 inches in diameter was substituted for the wire rope, $P - Q = 27$ lbs., of which 12 lbs. was due to the rigidity. Hence in one case the loss of work was less than $\frac{3}{4}$ of 1%, in the other about 1%, caused by the rigidity. For very small sheaves and thick ropes the loss is probably much greater. 19 194 MECHANICS OF ENGINEERING. 177. Miscellaneous Examples.—Example 1. The end of a shaft 13 inches in diameter and making 50 revolutions per minute exerts against its bearing an axial pressure of 10 tons and a lateral pressure of 40 tons. With $f = f' = 0.05$, required the H.P. lost in friction. **Ans.** 92.3 H. P. **Example 2.** A leather belt passes over a vertical pulley, covering half its circumference. One end is held by a spring balance, which reads 10 lbs. while the other end sustains a weight of 30 lbs., the pulley making 100 revolutions per minute. Required the coefficient of friction, and the H.P. spent in overcoming this friction. When the belt is turned in the other direction, the weight remaining the same. The diameter of the pulley is 18 inches. $$\text{Ans.} \quad f = 0.29 ; \quad f' = 0.149 \text{ and } 254 \text{ H. P.}$$ **Example 3.** A grindstone with a radius of gyration = 12 inches has been revolving at 120 revolutions per minute, and at a given instant is left to the influence of gravity and axle friction. The axle is 14 inches in diameter, and the wheel made 160 revolutions in coming to rest. Required the coefficient of axle-friction. **Ans.** $f = 0.389$. **Example 4.** A board $A$, weight 9 lbs., rests horizontally on another $B$; coefficient of friction of rest between them being $f = 0.30$. $B$ is now moved horizontally with a uniformly accelerated motion, the acceleration being = 15 feet per "square second" will $A$ keep company with it, or not? **Ans.** "No." A diagram showing a shaft with a pulley system and weights. PART III. STRENGTH OF MATERIALS. [OR MECHANICS OF MATERIALS]. CHAPTER I ELEMENTARY STRESSES AND STRAINS. 178. Deformation of Solid Bodies.—In the preceding portions of this work, what was called technically a "rigid body," was supposed incapable of changing its form, i.e., the positions of its particles relatively to each other, under the action of any forces to be brought upon it. This supposition was made because the change of form which must actually occur when applied to a solid body, such as straight lines, angles, etc., measured in any one body, among most of the pieces of a properly designed structure or machine. To show how the individual pieces of such constructions should be designed to avoid undesirable deformation or injury is the object of this division of Mechanics of Engineering, viz., the Strength of Materials. A diagram showing a circular cross-section with two parallel lines (A and B) and two perpendicular lines (C and D) intersecting at right angles. The lines A and B are labeled P and P respectively, and the lines C and D are labeled P and P respectively. Fig. 108. $178$ As perhaps the simplest instance of the deformation or distortion of a solid, let us consider the case of a prismatic rod in a state of tension. Fig. 192 (link of a surveyor's 196 MECHANICS OF ENGINEERING. chain, e.g.). The pull at each end is $P$, and the body is said to be under a tension of $P$ (i.e., tons, or other unit), not 2$P$. Let $ABCD$ be the end view of an elementary parallelopipedal element, i.e., a square prism with faces at 45° with the axis of the prism. It is now deformed, the four faces perpendicular to the paper being longer than before, while the angles $BAD$ and $BOD$, originally right angles, are now smaller by a certain amount $\Delta ABCD$ larger by an equal amount $\Delta$. The element is said to be in a state of strain, viz.: the elongation of its edges (parallel to paper) is called a tensile strain, while the alteration in the angles between its faces is called a shearing strain, angular distortion (sometimes also called a slanting, or tangential strain) $BCD$ being turned aside, relatively to $AD$, and thereby caused the change of angle. [This use of the word strain, to signify change of form and not the force producing it, is of recent adoption among many, though not all, technical writers.] 179. Strains. Two Kinds Only.—Just as a curved line may be considered to be made up of small straight-line elements, so the substance of any solid body may be considered to be made up of small contiguous parallelopipeds, whose angles are each 90° before the body is acted upon by forces but which are afterwards curvilinear. An line, or surface, forming for such a material is sometimes called a fibre, but this does not necessarily imply a fibrous nature in the material in question. The system of imaginary cutting surfaces by which the body is thus subdivided need not consist entirely of planes ; in the subject of Torsion, for instance, the parallelopipedal elements considered lie in concentric cylindrical shells, cut both by transverse and radial planes. Since these elements are too small to small that the only possible effect on them would be any one of them, as induced by a system of external forces acting on the body, are * When $\Delta$ is nearly 0 or 90°, $BC$ and $AD$ (or $AB$ and $DO$) are shorter than before, on account of lateral contraction; see § 193. **ELEMENTARY STRESSES, ETC.** 197 elongations or contractions of its edges, and alteration of its angles, there are but two kinds of strain, elongation (contraction, if negative) and shearing. 180. Distributed Forces or Stresses.—In the matter proceeding this chapter it has sufficed for practical purposes to consider a force as applied at a point of a body, but in reality it must be distributed over definite areas, for otherwise the material would be subjected to an infinite force per unit of area. (Forces like gravity, magnetic attraction, etc., we have already treated as distributed over the mass of a body, but reference is now had particularly to the pressure of one body against another, or the action of one portion of the body on the remainder.) For instance, sufficient surface must be provided between the end of a loaded beam and the pier on which it rests to avoid injury to the pier. Again, too great a tension may be used to sustain a given load, or the tension per unit of area of its cross section becomes sufficient to rupture it. *Stress is distributed force, and its intensity at any point of the area is* $$p = \frac{dF}{dF} \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1)$$ where $dF$ is a small area containing the point and $dP$ the force coming upon that area. If equal $dF's$ (all parallel) act on equal $dF's$ of a plane surface, the stress is said to be of uniform intensity, which is then $$p = P/F$$ \[ \text{where } P =\text{ total force and } F =\text{ the total area over which it acts. The steam pressure on a piston is an example of stress of uniform intensity.} \] (2) 198 MECHANICS OF ENGINEERING. For example, if a force $P = 28800$ lbs. is uniformly distributed over a plane area of $F = 72$ sq. inches, or $\frac{1}{2}$ of a sq. foot, the intensity of the stress is $$p = \frac{28800}{72} = 400 \text{ lbs. per sq. inch},$$ (or $p = 28800 + \frac{1}{2} = 57600$ lbs. per sq. foot, or $p = 14,400 + \frac{1}{2} = 28.8$ tons per sq. ft., etc.). 181. Stresses on an Element; of Two Kinds Only.—When a solid body of any material is in equilibrium under a system of forces which do not rupture it, not only is its shape altered (i.e. its elements are strained), and stresses produced on those elements on which the forces act, but other stresses are also produced on the surfaces between the spaces which separate element from element, (over and above the forces with which the elements may have acted on each other before the application of the external stresses or "applied forces"). So long as the whole solid is the "free body" under consideration, these internal stresses, being the forces with which the portion on one side of an imaginary cutting plane acts on the portion on the other side, do not appear in any equation of equilibrium (for if introduced they would cancel out). If we consider a free a portion only, that is all whose bounding surfaces are cutting planes of the original body, the stresses existing on these planes are brought into the equations of equilibrium. Similarly, if a single element of the body is treated by itself, the stresses on all six of its faces, together with its weight, form a balanced system of forces, the body being supposed at rest. A diagram showing a single element of a solid body with arrows indicating forces acting on it. Fig. 193 ELEMENTARY STRESSES, ETC. 199 As an example of internal stress, consider again the case of a rod in tension ; Fig. 193 shows the whole rod (or eye-bar) free, the force $P$ being the pressures of the pins in the eyes, and causing external stress (compression here) on the surfaces of contact. Conceive a right section made through $RS$, far enough from the eye, $C$, that we may consider the internal stresses to be uniform in this section, and consist of the stresses $BSO$ and $BSC$ shown in Fig. 193. The stresses on $RS$, now one of the bounding surfaces of the free body, must be parallel to $P$, i.e., normal to $RS$; (otherwise they would have components perpendicular to $P$, which is precluded by the necessity of $\Sigma Y$ being = 0, and the supposition of uniformity.) Let $F$ = the sec- Fig. 193. tional area $RS$, and $p$ = the stress per unit of area; then $$\Sigma X = 0 \text{ gives } P = Fp, \text{i.e., } p = \frac{P}{F}.$$ (2) The state of internal stress, then, is such that on planes perpendicular to the axis of the bar the stress is tensile and normal (to those planes). Since if a section were made oblique to the axis of the bar, the stress would still be parallel to the axis for reasons as above, it is evident that on an oblique section, the stress has components both normal and tangential to the section, the normal component being a tension. Fig. 196. 200 MECHANICS OF ENGINEERING. The presence of the tangential or shearing stress in oblique sections is rendered evident by considering that if an oblique dove-tail joint were cut in the rod, Fig. 195, the shearing friction on its surfaces may be sufficient to overcome friction and cause sliding along the oblique plane. If a short prismatic block is subjected to compressive action or tension, and applied centrally on one base, we may show that the state of internal stress is the same as that of the rod under tension, except that the normal stresses are of contrary sign, i.e., compressive instead of tensile, and that the shearing stresses (or tendency to slide) on oblique planes are opposite in direction to those in the rod. Since the resultant stress on a given internal plane of a body is fully represented by its normal and tangential components, we are therefore justified in considering but two kinds of internal stress, *normal* or direct, and *tangential* or shearing. **182 Stress on Oblique Section of Rod in Tension.—Consider free a small cubic element whose edges are $a$ in length; it has two faces parallel to the plane being taken near the middle of the rod in Fig. 192. Let the angle which the face $AB$, Fig. 196, makes with the axis of the rod be $a$. This angle, for our present purpose, is considered to remain the same while the two forces $P$ are acting, as before their action. The resultant force on this face $AB$, having an intensity $p = P + F$, (see eq. 2) per unit of transverse section of rod, is $= p (a \sin a)$ e. Hence its component normal to $AB$ is $$p^2 \sin^2 a;$$ and the tangential or shearing component along **ELEMENTARY STRESSES, ETC.** 201 $AB = p a^2 \sin a \cos a$. Dividing by the area, $a^2$, we have the following: For a rod in simple tension we have, on a plane making an angle, $a$, with the axis: a Normal Stress $\rightarrow p \sin^2 a$ per unit of area . (1) and a Shearing Stress $\rightarrow p \sin a \cos a$ per unit of area . (2) "Unit of area" here refers to the oblique plane in question, while $p$ denotes the normal stress, and $q$ the stress of area of a traverse section at right angles to 90°, Fig. 194. The stresses on $CD$ are the same in value as on $AB$, while for $BQ$ and $AD$ we substitute $90 - a$ for $a$. Fig. 197 shows these normal and shearing stresses, and also, much exaggerated, the strains or change of form of the element (see Fig. 192). 189a. Relation between Stress and Strain.—Experiment shows that so long as the stresses are of such moderate value that the piece recovers its original form completely when the external forces which induce the stresses are removed, the following is true and is known as Hook's Law (stress proportional to strain). As the forces $P$ in Fig. 193 (rod in tension) are gradually increased, the elongation, or additional length, of $BA$ increases in the same ratio as the normal stress, $p$, on the sections $RS$ and $KN$, per unit of length [§ 191]. As for the distorting effect of shearing stresses, consider in Fig. 197 that since $$p \sin a \cos a = p \cos(90^\circ - a) \sin(90^\circ - a)$$ the shearing stress per unit of area is of equal value as all four of the faces (perpendicular to paper) in the elementary block, and is evidently accountable for the shearing strain, i.e., for the angular distortion, or difference, $\delta$, between 90° and the present value of each of the four angles. According to Hook's Law then, as $P$ increases within the limit mentioned above, $\delta$ varies proportionally to $$p \sin a \cos a,$$ i.e., to the stress. A diagram showing a rod in tension with forces applied at different points along its length. 202 MECHANICS OF ENGINEERING. **182b Example.**—Supposing the rod in question were of a kind of wood in which a shearing stress of 300 lbs. per sq. inch along the grain, or a normal stress of 400 lbs. per sq. inch, perpendicular to a fibre-plane will produce rupture, required the value of the angle which the grain must make with the plane so that the maximum change of rupture from each source may be the same. This requires that $300 \cdot 400 \cdot p \sin a \cos x \cdot p \sin^2 a$, i.e. tan $a$ must = -2000 \cdot c = 63\frac{1}{2}^\circ$. If the cross section of the rod is 2 sq. inches, the force $P$ at each end necessary to produce rupture of either kind, when $a = 63\frac{1}{2}^\circ$, is found by putting $p \sin a \cos x_0 = 300 \cdot p = 500 \cdot 400$ lbs. per sq. inch. Whence, since $p = P/F$, $P = 1000$ lbs. (Units, inch and pound.) **183. Elasticity is the name given to the property which most materials have, to a certain extent, of regaining their original form when the external forces are removed. If the stress exceeds the elastic limit, called the Elastic Limit, the recovery of original form on the part of the elements is only partial, the permanent deformation being called the Set. Although theoretically the elastic limit is a perfectly definite stage of stress, experimentally it is somewhat indefinite, and is generally considered to be reached when the permanent set becomes well marked as the stresses are increased and the test piece is given ample time for recovery in the state of rest. The Safe Load or Stress, taken well within the elastic limit, determines the working strength or safe load of the piece under consideration. E.g., the tables of safe loads of the rolled wrought iron beams, for floors, of the New Jersey Steel and Iron Co., at Trenton, are computed on the theory that the greatest normal stress (tension or compression) occurring on any internal plane shall not exceed 12,000 lbs. per sq. inch; nor the greatest shearing stress 4,000 lbs. per sq. inch. A page from a book about mechanics of engineering. ELEMENTARY STRESSES, ETC. 203 The Ultimate Limit is reached when rupture occurs. 184. The Modulus of Elasticity (sometimes called co-efficient of elasticity) is the number obtained by dividing the stress per unit of area by the corresponding relative strain. Thus, a rod of wrought iron $\frac{3}{4}$ sq. inch sectional area being stretched to a tension of 5,500 lbs., i.e., it is found that a length which was 6 ft. before tension is $-6.002$ ft. during tension. The relative longitudinal strain or elongation is then $(0.002)+(-6)=1.3,000$ and the corres- ponding stress (being the normal stress on a transverse plane) has an intensity of $$p_1 = F = 5,000 \div \frac{3}{4} = 10,000 \text{ lbs.}, \text{ per sq. inch}.$$ Hence by definition the modulus of elasticity is (for ten- sion) $$E_1 = p_1 + (-\frac{1}{E_1}) = 30,000,000 \text{ lbs. per sq. inch}, (\text{the sub-script "t" refers to tension}).$$ It will be noticed that since $E$ is an abstract number, $E_1$ is of the same quality as $p_1$, i.e., lbs. per sq. inch, or one di- mension of force divided by two dimensions of length. (In the subject of strength of materials the inch is the most convenient English linear unit, when the pound is the unit of force; sometimes the foot and ton are used to- gether.) The foregoing would be called the modulus of elasticity of wrought iron in tension in the direction of the fibre, as given by the experiment quoted. But by Hooke's Law $p$ and $e$ vary together, for a given direction in a given ma- terial, hence within the elastic limit $E$ is constant for a given direction in a given material. Experiment confirms this approximately. Similarly, the modulus of elasticity for compression $E_c$, $$E_c = p_c + (-\frac{1}{E_c}) = 30,000,000 \text{ lbs. per sq. inch}, (\text{the sub-script "c" refers to compression}).$$ 204 MECHANICS OF ENGINEERING. in a given direction in a given material may be determined by experiments on short blocks, or on rods confined laterally to prevent flexure. As the modulus of elasticity for shearing, $E_s$, we divide the shearing stress per unit of area in the given direction by $\delta$ (in $\pi$ measure) the corresponding angular strain or distortion; e.g., for an angular distortion of 1° or $\delta = 0.0174$, and a shearing stress of 1,566 lbs. per sq. inch, we have $E_s = \frac{1566}{0.0174} = 9,000,000$ lbs. per sq. inch. Unless otherwise specified, by modulus of elasticity will be meant a value derived from experiments conducted within the elastic limit, and this, whether for normal stress or for shearing, is approximately constant for a given direction in a given substance.* 185. Isotropes.—This name is given to materials which are homogeneous as regards their elastic properties. In such a material the modulus of elasticity are individually the same for all directions. E.g., a rod of rubber cut out of a long cylinder has the same modulus of elasticity whether subjected to tension, whatever its original position in the mass. Fibrous materials like wood and wrought iron are not isotropic; the direction of grain in the former must always be considered. The "piling" and welding of numerous small pieces of iron prevent the resultant forging from being isotropic. 186. Resilience refers to the potential energy stored in a body held under external forces in a state of stress which does not pass the elastic limit. On its release from constraint, by virtue of its elasticity it can perform a certain amount of work called the resilience, depending in amount upon the circumstances of each case and the nature of the material. See § 145. 187. General Properties of Materials.—In view of some definitions already made we may say that a material is ductile *The modul, or "coefficients" of elasticity as used by physicists are well explained by Stewart and Dew's Practical Physics, Vol. I, pp. 104, etc. Their "coefficients of rigidity" are not so useful as those used by physicists. **ELEMENTARY STRESSES, ETC.** 205 when the ultimate limit is far removed from the elastic limit; that is it brittle like glass and cast iron, when those limits are put together. The same distinction of elasticity means that a comparatively small force is necessary to produce a given change of form, and vice versa, but implies little or nothing concerning the stress or strain at the elastic limit; thus Weibach gives $E$, lbs. per sq. inch for wrought iron $= 28,000,000$ double the $E$, for cast iron while the compressive stresses at the elastic limit are the same for both materials (nearly). 18th General Problem of Internal Stress—This, as treated in the mathematical Theory of Elasticity, developed by Lame, Clepoyson and Poisson, may be stated as follows: Given the original form of a body when free from stress; and certain co-efficients depending on its elastic properties; required the new position, the altered shape, and the intensity of the stress on each of the six faces, of every element of the body, when a given balanced system of forces is applied to the body. Solutions, by this theory, of certain problems of the nature just given involve elaborate, intricate, and bulky analysis; but for practical purposes Navier's theories (1838) and those of more recent date, are sufficiently exact, when their moduli are properly determined by experiments covering a wide range of cases and materials. These will be given in the present work, and are comparatively simple. In some cases graphic will be preferred to analytic methods as more simple and direct, and indeed for some problems they are the only methods yet discovered for obtaining solutions. Again, experiment is relied on almost exclusively in dealing with bodies of certain forms under peculiar systems of forces, empirical formulae being based on the experiments made; e.g., the collapsing of boiler flues, and in some degree the flexure of long columns. 206 MECHANICS OF ENGINEERING. 189. Classification of Cases.—Although in almost any case whatever of the deformation of a solid body by a balanced system of forces acting on it, normal and shearing stresses are both developed in every element which is affected at all (according to the plane section considered), still, cases where the body is prismatic, and the external system consists of two equal and opposite forces, one at each end of the piece and directed away from each other, are commonly called cases of Tension (Fig. 192); if the piece is a short prism with the same two principal forces directed toward each other, we have a case similar to the last, but where the prism is quite long ("long column"), is a case of Flexure or bending, as are also most cases where the "applied forces" (i.e., the external forces), are not directed along the axis of the piece. Riveted joints and "pin-connections" present cases of Shearing; a twisted shaft one of Torision. When the gravity forces due to the weights of the elements are also considered, a combination of two or more of the foregoing general cases may occur. In each case, as treated, the principal objects aimed at are, so to design the piece or its loading that the greatest stress, in whatever element it may occur, shall not exceed a safe value; and sometimes, furthermore, to prevent too great deformation on the part of the piece. The first object is to provide sufficient strength; the second sufficient stiffness. 190. Temperature Stresses.—If a piece is under such constraint that it is not free to change its form with changes of temperature, external forces are induced, the stresses produced by which are called temperature stresses. A diagram showing a prism with two applied forces at opposite ends. TENSION. 207 **TENSION.** 101. Hook's Law by Experiment.—As a typical experiment in the tension of a long rod of ductile metal such as wrought iron and the metal shown in the following table is quoted from Prof. Cotterill's "Applied Mechanics." The experiment is old, made by Hodgkinson for an English Railway Commission, but well adapted to the purpose. From the great length of the rod, which was of wrought iron and 0.517 in. in diameter, the portion whose elongation was observed being 49 ft. 8 in. long, the small increase in length below the elastic limit was readily measured. The successive loads were of such a value that the tensile stress $p = \frac{F}{A}$ remained constant throughout the test. The elongation of the same section, was made to increase by equal increments of 2057.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any.




















































































<
                                                                          etc.
P λ Δl e - λ + Δl λ'
Load (lbs. per square inch.) Elongation (inches.) Increment of Elongation (inches.) e, the relative elongation, i.e., $\frac{\Delta l}{\lambda}$ Permanent Set (inches.)
1000.5 0.049 0.049 0.00068
2x = 2001 0.061 0.061 0.00116
3x = 3001.5 0.066 0.066 0.00183
4x = 4002 0.068 0.068 0.00235
5x = 5002.5 0.072 0.072 0.00327
6x = 6003 0.076 0.076 0.00475
7x = 7003.5 0.081 0.081 0.00631
8x = 8004 0.085 0.085 0.01198
9x = 9004.5 0.119 0.119
A table showing elongations of a wrought iron rod under various loads. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of the same section, was made to increase by equal increments of 2957.5 lbs. per sq. in., its initial value. After each application of load the elongation was measured, and after the removal of the load, the permanent set, if any. A graph showing the relationship between load and elongation. The elongation of 208 MECHANICS OF ENGINEERING. Referring now to Fig. 198, the notation is evident. $F$ is the total load in any experiment, $F$ the cross section of the rod; hence the normal stress on the transverse section is $\sigma = \frac{P}{A}$. The loads are increased by equal increments, the corresponding increments of the elongation $\lambda$ should also be equal if Hooke's law is true. It will be noticed in the table that this is very nearly true up to the 8th loading, i.e., that $d\lambda$, the difference between two consecutive values of $\lambda$, is nearly constant. In other words the proposition holds good: $$P:F_1::\lambda_1:\lambda_2$$ If $P$ and $P_1$ are any two loads below the 8th, and $l$ and $l_1$ the corresponding elongations. The permanent set is just perceptible at the 3rd load, and increases rapidly after the 8th, as also the increment of elongation. Hence at the 8th load, which produces a tensile stress on the cross section of $p=8x267.5=21340.0$ lbs. per sq. inch, the elastic theory fails. As to the state of stress in the individual elements, if we conceive each sub-division of the rod that four edges of each element are parallel to the axis of the rod, we find that it is in equilibrium between two normal stresses on its end faces (Fig. 199) of a value $-\sigma dF_-$ and $+\sigma dF_+$ for the horizontal section of the element. If $dz$ was the original length, and $dl$ the elongation produced by $pdF$, we shall have, since all the $dz$'s of the length are equally elongated at the same time, $$\frac{dl}{dz} = \frac{1}{T}$$ Fig. 198 Fig. 199 TENSION. 209 where $l$ = total (original) length. But $\frac{dl}{dx}$ is the relative elongation $\epsilon$, and by definition (§ 184) the Modulus of Elasticity for Tension, $E = \frac{F}{A} = \frac{P}{A\cdot l}$. $\therefore E = \frac{dF}{dx}$ or $E = \frac{F}{dx}$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) Eq. (1) enables us to solve problems involving the elongation of a prism under tension, so long as the elastic limit is not surpassed. The values of $E$, computed from experiments like those just cited, would be the same for any load under the elastic limit, if Hooke's law were accurately obeyed, but in reality they differ somewhat, especially if the material lacks homogeneity. In the present instance (see Table) we have from the
3rd Exper.$E = p + t = 28,680,000$ lbs. per sq. in.
5th$E = -28,009,000$
8th$E = -27,848,000$
If similar computations were made beyond the elastic limit, i.e., beyond the 8th Exper., the result would be much smaller, showing the material to be yielding much more readily. 193. Strain Diagram—If we plot the stresses per sq. inch ($p$) as ordinates of a curve, and the corresponding relative elongations ($\epsilon$) as abscissas, we obtain a useful graphic representation of the results of experiment. Thus, the table of experiments just cited being utilized in this way, we obtain on paper a series of points through which a smooth curve may be drawn, viz.: OBC Fig. 200, for wrought iron. Any convenient scales may be used for $p$ and $\epsilon$; and experiments having been made on other metals in tension and the results plotted to the same scales 210 MECHANICS OF ENGINEERING. as before for $p$ and $\epsilon$, we have the means of comparing their tensile properties. Fig. 200 shows two other curves, representing (roughly) the average behavior of steel and cast iron. As regards the curve for $B$, $B'$, it will be noticed that the curve for wrought iron makes a sudden turn from the vertical, while those of the others curve away more gradually; that the curve for steel lies nearer the vertical axis than the others, which indicates a higher value for $E_1$; and that the ordinates $BA', BA''$, and $BA'''$ (respectively 31,000, 9,000, and 30,000 lbs. per sq. inch) indicate the tensile stress at the elastic limit. These latter quantities will be called the *moduli of tenacity at elastic limit* for the respective materials. [On a true scale the point $C$ would be much further to the right than here shown. Only one half of the curve for steel is given, for want of space.] Within the elastic limit the curves are nearly straight (proving Hooke's law) and if $\alpha$, $\alpha'$, and $\alpha''$ are the angles made by these straight portions with the axis of $X$ (i.e., of $\epsilon$), we shall have $$\text{(E. for w. iron):} E(\text{w. iron}) = E(\text{steel}): \tan \alpha : \tan \alpha' : \tan \alpha''$$ Fig. 200. 210 TENSION. 211 as a graphic relation between their moduli of elasticity (since $E = \frac{P}{A}$). Beyond the elastic limit the wrought iron rod shows large increments of elongation for small increments of stress, i.e., the curve becomes nearly parallel to the horizontal axis, until rupture occurs at a stress of 53,000 lbs. per sq. inch of original sectional area (at rupture this area is some- what reduced, especially in the immediate neighborhood of the section of rupture; see next article) and after a rela- tive elongation $\epsilon$ about 0.30 or 0.3%. (The preceding table shows only a portion of the results.) The curve for steel shows a much higher breaking stress, 100,000 lbs. per sq. inch, than wrought iron, but the relative elongation is smaller, $\epsilon$ about 10%. This is an average curve; tool steels give an elongation at rupture of about 4 to 5%, while soft steels resemble wrought iron in their ductility, giving an extreme elongation of from 10 to 20%. Their breaking stresses range from 70,000 to 150,000 lbs. or more per sq. inch. Cast iron, being comparatively brit- tle, reaches at rupture an elongation of only 3 or 4 tenths of one percent, the rupturing stress being about 18,000 lbs. per sq. inch. The elastic limit is near zero. In the case of this metal the proportion of carbon and the mode of manufacture have much influence on its be- havior under test. 193. Lateral Contraction.—In the stretching of prisms of nearly all kinds of material, accompanying the elongation of length is found also a diminution of width whose rela- tive amount in the case of the three metals just treated is about $\frac{1}{2}$ or $\frac{1}{3}$ of the relative elongation (within elastic limits). This lateral contraction was first observed by Hooke in 1679. This relative lateral contraction or decrease of diameter $\frac{1}{2}$ to $\frac{1}{3}$ of $e$, i.e., about 0.0008. In the case of cast iron and hard steels contraction is not noticeable ex- 212 MECHANICS OF ENGINEERING cept by very delicate measurements, both within and without the elastic limit; but the more ductile metals, as wrought iron and the soft steels, when stretched beyond the elastic limit show this feature of their deformation in a very marked degree. Fig. 201 shows by dotted lines the original contour of a wrought iron rod, while the continuous line represents that at which the condition of rupture, whose position is determined by some local weakness, the drawing out is peculiarly pronounced. The contraction of area thus produced is sometimes as great as 50 or 60% at the fracture. 194. Flow of Solids.—When the change in relative position of the elements of a solid is extreme, as occurs in the making of lead pipe, during which the material is drawn out to a ductile metal as in the preceding article, we have instances of what is called the "Flow of Solids," interesting experiments on which have been made by Tresca. 195. Moduli of Tenacity.—The tensile stress per square inch (of original sectional area) required to rupture a prism of a given material will be denoted by $T$ and called the modulus of ultimate tenacity; similarly, the modulus of safe tension will be denoted by $P$ and called the safe load; by $T'$; while the tensile stress at elastic limit may be called $T''$. The ratio of $T$ to $T'$ is not fixed in practice but depends upon circumstances (from $\frac{1}{3}$ to $\frac{2}{3}$). Hence, if a prism of any material sustains a total pull or load $P$, and has a sectional area $F$, we have $$P = FT' \quad \text{for the ultimate or breaking load}$$ $$P = FT'' \quad \text{safe load}$$ $$P = FT''' \quad \text{load at elastic limit}.$$ Of course $T$ should always be less than $T'$. (2) \begin{align*} & P = FT' \\ & P = FT'' \\ & P = FT''' \\ & T < T' \end{align*} 2 TENSION. 213 196. Resilience of a Stretched Prism—Fig. 302. In the gradual stretching of a prism, fixed at one extremity, the value of the tensile force $P$ at the other necessarily depends on the elongation $\lambda$ at each stage of the lengthening, according to the relation [eq. (1) of § 191.] $$\frac{P}{FE}, \quad \ldots \quad \ldots$$ (3) within the elastic limit. (If we place a weight $G$ on the flanges of the unstretched prism and then leave it to the action of gravity and the elastic action of the prism, the weight begins to sink, meeting an increasing pressure $P$, proportional to $\lambda$, from the flanges). Suppose the stretching to continue until $P$ reaches some value $P'$ (at elastic limit say), and $\lambda$ a value $\lambda'$. Then the work done so far is $$U = \text{mean force} \times \text{space} = \frac{1}{2} P' \lambda'$$ (4) But from (2) $F = PT'$, and (see §§ 184 and 191) $$\lambda' = e^{\gamma}$$ :. (4) becomes $U = \frac{1}{2} T' e^{\gamma} F' = \frac{1}{2} T' e^{\gamma} F$$ (5) where $F$ is the volume of the prism. The quantity $\frac{1}{2} T' e^{\gamma}$, or work done in stretching to the elastic limit a cubic inch (or other unit of volume) of the given material, Weibach calls the *Modulus of Resilience for tension*. From (5) it appears that the amounts of work done in stretching to the elastic limit prisms of the same material but of different dimensions are proportional to their volumes simply. The quantity $\frac{1}{2} T' e^{\gamma}$ is graphically represented by the area under this curve for a prism $OAB$, O'B', etc., Fig. 300; for (in the curve for wrought iron for instance) the modulus of tenacity at elastic limit is represented by $AB$, and $\epsilon$ (i.e., $\epsilon$ for elastic limit) by $OA$. The remainder of A diagram showing a stretched prism with a weight on its flanges. 302 214 MECHANICS OF ENGINEERING. the area $OBC$ included between the curve and the horizontal axis, i.e., from $B$ to $C$, represents the work done in stretching a cubic unit from the elastic limit to the point of rupture, for each vertical strip having an altitude $-p$ and a width $-d\epsilon$, has an area $-pd\epsilon$, i.e., the work done by the stress $p$ on one face of a cubic unit through the distance $d\epsilon$, or increment of elongation. If a weight or load $G$ be "suddenly" applied to stretch the prism, i.e., placed on the flanges, barely touching them, and then allowed to fall, when it comes to rest again it has passed through the elastic limit and experienced at this instant some pressure $P_1$ from the flanges, $P_1 = G$. The work $W_1$ has been entirely expended in stretching the prism, none in changing the kinetic energy of $G$, which was zero at both beginning and end of the distance $\lambda_1$, $$G\lambda_1 = \frac{1}{2}P_1\lambda_1$$ $$P_1 = 2G.$$ Since $P_1 = 2G$, i.e., is > $G$, the weight does not remain in this position but is pulled upward by the elasticity of the prism. In fact, the motion is harmonic (see §§ 59 and 138). Theoretically, the elastic limit not being passed, the oscillations should continue indefinitely. Hence a load $G$ "suddenly applied" occasions double the tension it would if compelled to sink gradually by a support underneath, which is not removed until the tension is just equal to $G$, causing no further prevention. If the weight $G$ sinks through a distance $\lambda_2 = h$ before striking the flanges, Fig. 202, we shall have similarly, within elastic limit, if $\lambda_2$ greatest elongation, (the mass of rod being small compared with that of $G$). $$G(h + \lambda_2) = \frac{1}{2}P_2\lambda_2$$ (6) If the elastic limit is to be just reached we have from eqa. (5) and (6), neglecting $\lambda_2$ compared with $h$, $$Gh = \frac{1}{2}T^* \cdot V$$ (7) TENSION. 215 an equation of condition that the prism shall not be injured. **Example.**—If a steel prism have a sectional area of $\frac{1}{4}$ sq. inch and a length $=10$ ft. =120 inches, what is the greatest allowable height of fall of a weight of 200 lbs., that the final tensile stress induced may not exceed $T=$ 30,000 lbs. per sq. inch, if $e'=.002$? From (7), using the inch and pound, we have $$h = \frac{T \cdot e' \cdot F}{2G} = \frac{30,000 \times .002 \times \frac{1}{4} \times 120}{2 \times 200} = 4.5 \text{ inches}.$$ 197. Stretching of a Prism by Its Own Weight.—In the case of a very long prism such as a mining-pump rod, its weight must be taken into account as well as that of the terminal load. In Fig. 186, (a) shows the prism as shown in its unstrained condition; at (b) strained by the load $P$, and its own weight. Let the cross section be $F$, the heaviness of the prism $=y$. Then the relative extension of any element at a distance $x$ from o is $$\epsilon = \frac{dL}{dx} = \frac{(P + yFy)}{FE}. \quad (1)$$ (See eq. (1) § 191); since $P_+ Fy$ is the load hanging upon the cross section at that locality. Equal $dL$s, therefore, are unequally elongated, $z$ varying from 0 to $l$. The total elongation is $$\lambda = \int_0^l dx = \frac{1}{FE}\int_0^l (P dx + y F dy) = \frac{P l}{FE} + \frac{yGl}{FE}.$$ La, $\lambda$ is the amount due to $P$, plus an extension which half the weight of the prism would produce, hung at the lower extremity. 216 MECHANICS OF ENGINEERING. The foregoing relates to the deformation of the piece, and is therefore a problem of stiffness. As to the strength $T$ of the prism, the relative elongation $\epsilon = \frac{dL}{dx}$ [see eq. (1)], which is variable, must nowhere exceed a safe value $r = T + E$, (from eq. (1) § 191, putting $P = FT$, and $i = x$). Now the greatest value of the ratio $\frac{dL}{dx}$, by inspecting eq. (1), is seen to be at the upper end where $x = L$. The proper cross section $F$, for a given load $P$, is thus found. Putting $\frac{P}{FE} = \frac{T}{E}$, we have $F = \frac{P}{T - rL}$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2) 198. Solid of Uniform Strength in Tension, or hanging body of minimum material supporting its own weight and a terminal load $P$. Let it be a solid of revolution. If every cross-section $F$ at a distance $x$ from the lower extremity, bears its safe load $FT$, every element of this solid will be in equilibrium in the form is the most economical of material. The lowest section must have an area $F_0 = P_0 + T$, since $P_0$ is its safe load. Fig. 204. Consider any horizontal lamina; its weight is $rFdz$, ($r$ = heaviness of the material, supposed homogeneous), and its lower base $F$ must have $P_0 + G$ for its safe load, i.e. $$G + P_0 = FT$$ (1) in which $G$ denotes the weight of the portion of the solid below $F$. Similarly for the upper base $F + dF$, we have $$G + P_0 + dFdz = (F + dF)T$$ (2) By subtraction we obtain $$rFdz = T'dF; \text{ i.e. } \frac{L}{T}dz = \frac{dF}{F}$$ TENSION. 217 in which the two variables $x$ and $F$ are separated. By integration we now have $$\int_{0}^{x} \frac{dF}{F} dx = \int_{r_0}^{r} \frac{dF}{F}; \text{ or } \frac{x}{T} = \log_{e} \frac{F}{F_0} \dots (3)$$ i.e., $F = F e^{\frac{x}{T}} = P_1 e^{\frac{x}{T}} \dots (4)$ from which $F$ may be computed for any value of $x$. The weight of the portion below any $F$ is found from (1) and (4); i.e. $$G = P_1 (e^{\frac{x}{T}} - 1) \dots (5)$$ while the total extension $l$ will be $$l = x' \frac{T'}{T'} l \dots (6)$$ the relative elongation $\frac{dl}{dx}$ being the same for every $dx$ and bearing the same ratio to $x'$ (at elastic limit), as $F'$ does to $F''$. 190. Tensile Stresses Induced by Temperature.—If the two ends of a prism are immovably fixed, when under no strain and at a temperature $t$, and the temperature is then lowered to a value $\ell$, the body suffers a tension proportional to the fall in temperature (within elastic limit). If for a rise or fall of 1° Fahr. (or Cent.) a unit of length of the material would change in length by an amount $\gamma$ (called the co-efficient of expansion), a rod whose ends would be constricted an amount $i = \gamma(t - t')$ during this given fall of temperature if one end were free. Hence, if this contraction is prevented by fixing both ends, the rod must be under a tension $P$, equal in value to the force which would be A crescent moon with a small star-like figure on top. 218 MECHANICS OF ENGINEERING. necessary to produce the elongation $i$, just stated, under ordinary circumstances at the lower temperature. From eq. (1) §191, therefore, we have for this tension due to fall of temperature $$P = \frac{E.F}{l} \eta(t-t_0) = E.F(t-t_0)\eta$$ For 1° Cent. we may write
For Cast iron $\eta = 0.000111$;
" Wrought iron" $= 0.000130$;
" Steel" $= 0.000108$ to $0.000114$;
" Copper" $\eta = 0.000172$;
" Zinc" $\eta = 0.000300$.
COMPRESSION OF SHORT BLOCKS. 200. Short and Long Columns.—In a prism in tension, its own weight being neglected, all the elements between the locality of application of the two external forces pro- ducing the stretching are in the same state of stress if the external forces act axially (excepting the few elements in the immediate neighborhood of the forces; these suffering local stresses dependent on the manner of application of the external forces), and the prism may be of any length without vitiating this statement. But if the two external forces are directed toward each other the intervening ele- ments will not all be in the same state of compressive stress unless all the prism is comparatively short (or unless numerous points of lateral support are provided). A long prism is thus subject to bending, thus even inducing tensile stress, in some cases, in the convex side. Hence the distinction between short blocks and long columns. Under compression the former yield by crushing or splitting, while the latter give way by flexure (i.e. bending). Long columns, then will be treated separately COMPRESSION OF SHORT BLOCKS. 219 in a subsequent chapter. In the present section the blocks treated being about three or four times as long as wide, all the elements will be considered as being under equal compressive stresses at the same time. 201. Notation for Compression.—By using a subscript $c$, we may write $E_c = \text{Modulus of Elasticity};^*$ i.e. the quotient of the compressive stress per unit of area divided by the relative shortening; also $C = \text{Modulus of crushing};$ i.e. the force per unit of sectional area necessary to rupture the block by crushing; $C' = \text{Modulus of safe compression},$ a safe compressive stress per unit of area; and $C'' = \text{Modulus of compression at elastic limit}.$ For the absolute and relative shortening in length we may still use $\lambda$ and $\epsilon$, respectively, and within the elastic limit may write equations similar to those for tension, $F$ being the sectional area of the block and $P$ one of the terminal forces, while $p = \text{compressive stress per unit of area of } F,$ viz: $$E_c = \frac{P}{\epsilon} = \frac{P + F}{\lambda + d\lambda} = \frac{P + F}{\lambda + d\lambda} = \frac{Pl}{Fl} \quad . . . . (1)$$ within the elastic limit. Also for a short block Crushing force $= FC$ Compressive force at elastic limit $= FC''$ Safe compressive force $= FC'$ (2) 202. Remarks on Crushing.—As in § 182 for a tensile stress, so for a compressive stress we may prove that a *Note.—It must be remembered that the modulus of elasticity, whether for normal or shearing stress, is a number indicative of stiff- ness, not strength. It is only by experiment (except that experiments to determine it must not pass that limit.) 220 MECHANICS OF ENGINEERING. shearing stress = p sin a cos a is produced on planes at an angle $a$ with the axis of the short block, $p$ being the compression per unit of area of transverse section. Accordingly it is found that short blocks of many comparatively brittle materials yield by shearing on planes making an angle of about 45° with the axis, the expression $p \sin a \cos a$ reaching a maximum for $\alpha = 45°$, that is wedge-shaped pieces form under these conditions. Hence the necessity of making the block three or four times as long as wide, since otherwise the friction on the ends would cause the piece to show a greater resistance by hindering this lateral motion. Crushing by splitting into pieces parallel to the axis sometimes occurs. Blocks of ductile material, however, yield by swelling out, or bulging, laterally, resembling plastic bodies some-what in their behavior. The elastic limit is more difficult to locate than in ten- sion, but seems to have a position corresponding to that in tension, in the case of wrought iron and steel. With cast iron, however, the relative compression at elastic limit is about double the relative extension (at elastic limit in tension), but the force producing it is also double. For all three metals it is found that $E_e = E_t$ quite nearly, so that the single symbol $E$ may be used for both. EXAMPLES IN TENSION AND COMPRESSION. 203. Tables for Tension and Compression.—The round num- bers of the following tables are to be taken as rude averages only, for use in the numerical examples following. (The scope and design of the present work admit of nothing more. For abundant detail of the results of the more im- portant experiments of late years, the student is referred to the recent works of Profs. Thurston, Burr, Lanza, and Wood.) Another column might have been added giving the Modulus of Resilience in each case, viz.: $\frac{1}{2} c^T$ (which also = $\frac{1}{2L_0}$); see § 196. $c$ is an abstract num- A page from a book discussing mechanics of engineering. EXAMPLES IN TENSION AND COMPRESSION. 221 ber, and =l+l, while $E$, $T$, and $T$ are given in pounds per square inch:
TABLE OF THE MODULI, ETC., OF MATERIALS IN TENSION.
Material $E$ $E$ at rupture Mod. of Elast. Elastic Limit $T$1st $T$2nd $T$3rd
abt. number abt. number lb. per sq. in. lb. per sq. in. lb. per sq. in. lb. per sq. in. lb. per sq. in.
Soft Steel, .00000 .0000 80,000,000 55,000 80,000
Hard Steel, .00000 .0000 40,000,000 97,000 190,000
Cast Iron, .00098 .0988 14,000,000 5,000 14,000
Weir Iron, .00986 .9865 28,000,000 21,926 21,926
To be used with the following values:
Glass,.19356.19356
Wood, with the three:.19356.19356
Bump rope,.19356.19356
                                                             7.0 <N.B.>--Expressed in kilograms per square centim., $E$, $T$ and $T$2nd would be numerically about $\frac{1}{4}$ as large as above, while $e$ and $e^{\prime}$ would be unchanged.
TABLE OF MODULI, ETC.; COMPRESSION OF SHORT BLOCKS.
Material $e$' $e$ at rupture Mod. of Elast. Elastic Limit $C$'
abt. number abt. number
lb. per sq. in.
                          &nb... 222 MECHANICS OF ENGINEERING 204. Example No. 1.—A bar of tool steel, of sectional area $=0.097$ sq. inches, is ruptured by a tensile force of 14,000 lbs. A portion of its length, originally $\frac{1}{2}$ foot, is now found to have a length of 0.532 ft. Required $T$, and $\epsilon$ at rupture. Using the inch and pound as units (as in the foregoing table) we have $T = \frac{144328}{(0.532 \times 12)} = 0.664$ lbs. per sq. in.; (eq. (2) § 195); while $$\epsilon = \left(\frac{0.532 - 0.5}{0.5} \times 12 + (0.5 \times 12)\right) = 0.064.$$ Example 2.—Tensile test of a bar of "Hay Steel" for the Glasgow Bridge, Missouri. The portion measured was originally 3.21 ft. long and 2.09 in. × 1.10 in. in section. At the elastic limit $P$ was 134,400 lbs., and the elongation was 0.064 ins. Required $E$, $T'$, and $\epsilon'$ (for elastic limit). $$\epsilon' = \frac{\lambda}{l} = \frac{0.064}{3.21 \times 12} = 0.00165 \text{ at elastic limit}.$$ $$T' = 134,400 \times (2.09 \times 1.10) = 54,000 \text{ lbs. per sq. in}.$$ $$E = \frac{P}{\epsilon'} = \frac{134,400}{0.00165} = 82,570,000 \text{ lbs. per sq. in}.$$ Nearly the same result for $E$ would probably have been obtained for values of $p$ and below the elastic limit. The Modulus of Resilience of the above steel (see § 196) would be $\frac{1}{2} e^{' } T' = 44.82$ inch-pounds per sq. in., per cubic inch of metal, so that the whole work expended in stretching to the elastic limit the portion above cited is $$U = \frac{1}{2} e^{' } T' V' = 3968 \text{ inch-lbs}.$$ An equal amount of work will be done by the rod in recovering its original length. Example 3.—A hard steel rod of $\frac{1}{4}$ sq. in. section and 20 ft. long is under no stress at a temperature of $130^\circ$ A bar of tool steel with a tensile test apparatus attached. EXAMPLES IN TENSION AND COMPRESSION. 223 Cent., and is provided with flanges so that the slightest contraction of length will tend to bring two walls nearer together. If the resistance to this motion is 10 tons how low must the temperature fall to cause this motion? $\gamma$ being $=.000129$ (Cent. scale). From § 199 we have, expressing $P$ in lbs. and $F$ in sq. inches, since $E_0=40,000,000$ lbs. per sq. inch, $$10 \times 2,000 = 40,000,000 \times \frac{1}{2} \times (130 - c) \times 0.000012; \text{ whence } c = 46.5^\circ\text{ Centigrade}.$$ **EXAMPLE 4.—If the ends of an iron beam bearing 5 tons at its middle rest upon stone piers, required the necessary bearing surface at each pier, putting $C'$ for stone $= 200$ lbs. per sq. inch. **EXAMPLE 5.—How long must a wrought iron rod be, supported vertically at its upper end, to break with its own weight? ** EXAMPLE 6.—One vesseoir (or boiler) was being pressed its neighbor with a force of 50 tons, the joint having a surface of 5 sq. feet; required the compression per sq. inch. 138.8 lbs. per sq. in., Ans. 205. Factor of Safety.—When, as in the case of stone, the value of the stress at the elastic limit is of very uncertain determination by experiment, it is customary to refer the value of the safe stress to that of the material, taking it the $n$'th portion of the latter. $n$ is called a factor of safety, and should be taken large enough to make the safe stress come within the elastic limit. For stone, $n$ should not be less than 10, i.e. $C' = C + n$; (see Ex. 6, just given). 206. Practical Notes.—It was discovered independently by Commander Beardale and Prof. Thurston, in 1873, that if wrought iron rods were strained considerably beyond the elastic limit and allowed to remain free from stress 224 MECHANICS OF ENGINEERING. for at least one day thereafter, a second test would show higher limits both elastic and ultimate. When articles of cast iron are imbedded in oxide of iron and subjected to a red heat for some days, the metal loses most of its carbon, and is thus nearly converted into wrought iron, lacking, however, the property of welding. Being malleable, it is called malleable cast iron. **Chromestel (iron and chromium) and tungsten steel pos- sess peculiar hardness, fitting them for cutting tools, rock drills, picks, etc.** By experiments on metals we understand the fact, recently dis- covered by Wöhler in experiments made for the Prussian Government, that rupture may be produced by causing the stress on the elements to vary repeatedly between two limiting values, the highest of which may be considerably below $T$ (or $C$), the number of repetitions necessary to produce rupture being dependent both on the range of variation and the higher value. For example, in the case of Phönix iron in tension, rupture was produced by causing the stress to vary from 0 to 58,800 lbs. per sq. inch 600 times, also from 0 to 44,900 lbs. per sq. inch 340,857 times; while 4,000,000 va- riations between 26,400 and 48,400 per sq. inch did not cause rupture. Many other experiments were made and the following conclusions drawn (among others): Unlimited repetitions of variations of stress (lbs. per sq. in.) between the limits given below will not injure the metal (Prof. Burr's Materials of Engineering.)
From 17,600 Comp. to 17,600 Tension From 30,800 Comp. to 30,800 Tension
Wrought Iron. " 0 " 33,000 "
Axle Cast Steel. " 0 " 52,900 "
" 88,800 Tens. to 88,000 "
(See p. 233 for an addendum to this paragraph.) SHEARING. 225 SHEARING. 207. Rivets.—The angular distortion called shearing strain in the elements of a body, is specially to be provided for in the case of rivets joining two or more plates. This distortion is shown, in Figs. 206 and 206, in the elements near the plane of contact of the plates, much exaggerated. A diagram showing the shearing strain in a rivet joint. Fig. 206 In Fig. 206 (a lap-joint) the rivet is said to be in single shear; in Fig. 206 in double shear. If $P$ is just great enough to shear off the rivet, the modulus of ultimate shear- ing, which may be denoted by $S$, is the shearing force per unit of section when rupture occurs) $$S = \frac{P}{F} = \frac{P}{\pi d^2} \quad \ldots \quad \ldots \quad \ldots \quad (1)$$ in which $F$=the cross section of the rivet, its diameter being $=d$. For safety a value $S = \frac{1}{4}$ to $\frac{1}{3}$ of $S$ should be taken for metal, in order to be within the elastic limit. As the width of the plate is diminished by the rivet hole the shearing force increases, and it becomes necessary to sustain the tension $P_1$ or $2P$, (according to the plate considered, see Fig. 206), $P$ being the safe shearing force for the rivet. Also the thickness $t$ of the plate should be such that the side of the hole shall be secure against crushing; $P$ must not be > $Ctel$, Fig. 205. Again, the distance $a$, Fig. 205, should be such as to prevent the tearing or shearing out of the part of the plate between the rivet and edge of the plate. 226 MECHANICS OF ENGINEERING. For economy of material the seam or joint should be no more liable to rupture by one than by another, of the Figure 307. four modes just mentioned. The relations which must then subsist will be illustrated in the case of the "butt-joint" with two cover-plates, Fig. 207. Let the dimensions be denoted as in the figure and the total tensile force on the joint be $Q$. Each rivet (see also Fig. 205) is exposed in each of two of its sections to a shear of $\frac{1}{2} Q$, hence for safety against shearing of rivets we put $$\frac{1}{2} Q = \frac{1}{2} \pi d^2 S$$ (1) Along one row of rivets in the main plate the sectional area for resisting tension is reduced to $(b - 3d)t$, hence for safety against rupture of that plate by the tension $Q$, we put $$Q = (b - 3d)tT'$$ (2) Equations (1) and (2) suffice to determine $d$ for the rivets and $t_t$ for the main plates, $Q$ and $b$ being given; but the values thus obtained should also be examined with reference to the compression in the side of the rivet hole, i.e., $\frac{1}{4} Q$ must not be > $C't_d$. [The distance $a$, Fig. 205, to the edge of the plate is recommended by different authorities to be from $d$ to $3d$.] Similarly, for the cover-plate we must have $$\frac{1}{2} Q \text{ or } (b - 3d)tT'$$ (3) and $\frac{1}{4} Q$ not > $C't_d$ and SHEARING. 227 If the rivets do not fit their holes closely, a large margin should be allowed in practice. Again, in boiler work, the pitch, or distance between centers of two consecutive rivets may need to be smaller, to make the joint steam-tight, than would be required for strength alone. 208. Shearing Distortion.—The change of form in an element due to shearing is an angular deformation and will be measured in $x$-measure. This angular change or difference between the value of the corner angle during strain and $\frac{1}{2}\pi$, its value before strain, will be called $\delta$, and is proportional (within elastic limits) to the shearing stress per unit area. In Fig. 208, all the four faces whose angles with each other have been changed. Fig. 208. (See § 181). By § 184 the Modulus of Shearing Elasticity is the quotient obtained by dividing $p_x$ by $\delta$; i.e., (elastic limit not passed), $$E_x = \frac{p_x}{\delta} \quad \ldots \quad \ldots \quad (1)$$ or inversely, $$\delta = p_x + E_x \quad \ldots \quad \ldots \quad (1')$$ The value of $E_x$ for different substances is most easily determined by experiments on torsion in which shearing is the most prominent stress. (This prominence depends on the position of the bounding planes of the element considered; e.g., in Fig. 208, if another element were considered within the one there shown and with its planes at 45° with those of the first, we should find tension alone on one pair of opposite faces, compression alone on the other pair.) It will be noted that shearing stress cannot be present on two opposite faces only, but exists also on another pair of faces (those perpendicular to the stress on the first), forming a couple of equal and opposite moment to the first, this being necessary for the equilibrium of the element, even when A diagram showing a rectangular element with a diagonal line indicating shearing distortion. The top face is rotated slightly clockwise relative to the bottom face. 228 MECHANICS OF ENGINEERING. Tensile or compressive stresses are also present on the faces considered. 208. Shearing Stress is Always of the Same Intensity on the Four Faces of an Element. (By intensity is meant per unit of area; and the four faces referred to are those perpendicular to the paper in Fig. 208, the shearing stress being parallel to the paper.) Let $ds$ and $dy$ be the width and height of the element in Fig. 208, while $dp$ is its thickness perpendicular to the paper. Let the intensity of the shear on the right hand face be $-q$, that on the top face $=p$. Then for the element as a free body, taking moments about the axis $O$ perpendicular to paper, we have $$q \cdot ds \cdot dy \times ds - p \cdot ds \cdot dy \times ds = 0 : q_s = p$$ ($dx$ and $dz$ being the respective lever arms of the forces $q$, $ds \cdot dy$ and $p$, $ds \cdot dy$.) Even if there were also tensions (or compressions) on one or both pairs of faces their moments about $O$ would balance (or fail to do so by a differential of a higher order) independently of the shears, and the above result would still hold. 210. Table of Moduli for Shearing.
Material $E^{\prime}$ $B^{\prime\prime}$ $S$
Modulus of Elasticity (Elastic limit.) (Resonance)
arc in measure Im. per sq. in. Im. per sq. in. Im. per sq. in.
Soft Steel, 9,000,000 70,000
Hard Steel, 9,000 45,000 95,000
Cast Iron, 3,001 7,000,000 15,000 30,000
Wrought Iron, 3,008 6,000,000 95,000 95,000
Bronze, 3,001
Glass,
Wood, screws,
Fibre,
Wood, along fibre,
Wood, screws, Fibre, Wood, along fibre, (1) (2) (3) (4) SHEARING. 229 As in the tables for tension and compression, the above values are averages. The true values may differ from these as much as 30 per cent. in particular cases, according to the quality of the specimen. 211. Punching rivet holes in plates of metal requires the overcoming of the shearing resistance along the convex surface of the cylinder punched out. Hence if $d$ = diameter of hole, and $t$ = the thickness of the plate, the necessary force for the punching, the surface sheared being $F = txd$, is $$P = S \times d$$ (2) Another example of shearing action is the "stripping" of the threads of a screw, when the nut is forced off longitudinally without turning, and resembles punching in its nature. 212. $E$ and $R$: Theoretical Relation.—In case a rod is in tension within its elastic limit, the relative times of lateral contraction (let this one) is no connected with $E$, and that if two of the three are known the third can be deduced theoretically. This relation is proved as follows, by Prof. Burr. Taking an elemental cube with four of its faces at 45° with the axis of the piece, Fig. 209, the axial half-diagonal $AD$ becomes of length $AD = mAD + AD$ under stress, while the transverse half diagonal contracts to a length $BD = AD - mAD$. The angular distortion $\delta$ A diagram showing a cube with faces at 45° angles to the axis, labeled A, B, C, D. Fig. 208. § 212. A diagram showing a triangle with sides labeled E, F, G, H. Fig. 209. § 212. The angular distortion $\delta$ 230 MECHANICS OF ENGINEERING. is supposed very small compared with 90° and is due to the shear $p_s$ per unit of area on the face $BC$ (or $BA$). From the figure we have $$\tan(45^\circ - \frac{\delta}{2}) = \frac{BD'}{AD} = 1 - m - e, \text{ approx.}$$ [But, Fig. 210, $\tan(45^\circ - x) = 1 - 2x$ nearly, where $x$ is a small angle, for, taking $\overline{CD} = unity = AE$, $\tan AD = AF = AE - EF$. Now approximately $EF = EA \cdot \sqrt{2}$ and $EG = BD \cdot \sqrt{2} = x \cdot \sqrt{2}$. $AF = 1 - 2x$ nearly.] Hence $$1 - \delta = 1 - m - e; \text{ or } \delta = m + e.$$ (2) Eq. (2) holds good whatever the stresses producing the deformation, but in the present case of a rod in tension, if it is an isotrope, and if $p_s$ = tension per unit of area on its transverse section, (see § 182, putting $a = 45^\circ$), we have $E_s = p_s + e$ and $E_s = p_s$ on $BC + d = \frac{1}{2}p_s + d$. Putting also $(m : e) = r$, whence $m = rt$, eq. (2) may finally be written $$\frac{1}{2E_r} = (r + 1)\frac{1}{E_r}; \text{ i.e., } E_r = \frac{E_s}{2(1 + r)}.$$ (3) Prof. Bauschinger, experimenting with cast iron rods, found that in tension the ratio $m : e$ was $\frac{m}{e}$ as an average, which in eq. (3) gives $$E_r = \frac{100}{240} E_s = \frac{5}{6} E_s \text{ nearly.}$$ (4) His experiments on the torsion of cast iron rods gave $E_s = 6,000,000$ to $7,000,000$ lbs. per sq. inch. By (4), then, $E_r$ should be $15,000,000$ to $17,500,000$ which is approximately true (§ 183). Corresponding results may be obtained for short blocks in compression, the lateral change being a dilatation instead of a contraction. SHEARING. 231 §13. Examples in Shearing. **EXAMPLE 1.**—Required the proper length, $a$, Fig. 311, to guard against the shearing off, along the grain, of the portion $ab$, of a wooden tie-rod, the force $P$ being = 2 tons, and the width of the tie = 4 inches. Using a value of $S' = 100$ lbs. per sq. in., we put $6 \times S' = 4,000$ cos 45°; i.e., $d = (4,000 \times \sqrt{2}) / (4 \times 100) = 7.07$ inches. **EXAMPLE 2.**—A $\frac{1}{4}$ in. rivet of wrought iron, in single shear (see Fig. 305) has an ultimate shearing strength $F' = F \cdot S' = F \cdot (\frac{1}{4} \pi d^2 / 4) \times 50,000 = 30,050$ lbs. For safety, putting $S' = 8,000$ instead of $S$, $F' = 4,800$ lbs. is its safe shearing strength in single shear. The wrought iron plate, to be secure against the side-crushing in the hole, should have a thickness $t$, computed thus: $$F' = tdC ; \text{or } 4,800 = t \cdot \frac{1}{4} \pi \times 12,000 : t = 0.46 \text{ in}.$$ If the plate were only 0.23 in. thick the safe value of $P$ would be only $\frac{1}{4} \pi$ of 4,800. **EXAMPLE 3.**—Conversely, given a lap-joint, Fig. 205, in which the plates are $\frac{1}{4}$ in. thick and the tensile force on the joint is 600 lbs. per linear inch of seam, how closely must $\frac{1}{4}$ inch rivets be spaced in one row, putting $S' = 8,000$ and $C = 12,000$ lbs. per sq. in.? Let the distance between centres of rivets be $= x$ (in inches), then the force upon each rivet = 600x, while its section $F = 0.44$ sq. in. Having regard to the shearing strength of the rivet we put $600x = 0.44 \times 8,000$ and obtain $x = 5.86$ in.; but considering that the safe crushing resistance of the hole is $\frac{1}{4} \pi \cdot \frac{1}{4} \times 12,000 = 2525$ lbs., this gives a pitch which is too large; what is the pitch to be adopted so that the tensile strength of the reduced sectional area of the plate, with this pitch? 232 MECHANICS OF ENGINEERING. **EXAMPLE 4.—Double butt-joint; (see Fig. 207): ¼ inch plate; $t_1$ in. rivets; $T = O - 12,000$; $S = 8,333$; width of plates=14 inches. Will one row of rivets be sufficient at each side of joint, if $Q = 30,000$ lbs.? The number of rivets = ? Here each rivet is in double shear and has therefore a double strength as regards shear. In double shear the safe strength of each rivet = $2FS = 7,333$ lbs. Now $30,000 + 7,333 = 40.8$ (say). With the four rivets in one row the re- duced sectional area of the main plate is $(14 - 4) \times 1 \times 1 \times 4 = 4.12$ sq. in., whose tensile test strength is $F'T = 6,119 \times 12,000 = 49,440$ lbs.; which is a $30,000$ lbs. r. main plate is safe in this respect. But as to side-crushing in holes in main plate we find that $C'_{\text{td}}$ (i.e. $12,000 \times 4 \times 1 \times 4 = 3,375$ lbs.) is $4$ i.e. $7,500$ lbs., the actual force on side of hole. Hence four rivets in one row are too few unless thickness of main plate be doubled. Will eight in one row be safe? Sibbald's Addendum to § 306. Elasticity of Stone and Cementa. Experiments by Gen. Gillmore with the large Watertown testing-machine in 1888 resulted as follows (see p. 221 for notation): With cubes of Haverstraw Freestone (a homogeneous brown- stone) from 1 in. to 12 in. on the edge, $E_c$ was found to be from 900,000 to 1,000,000 lbs. per sq. in. approximately; and $C$ about 4,000 or 5,000 lbs. per sq. in. Cubes of the same range of sizes as Dyckerman's Portland cement gave $E_c$ from 1,555,000 to 1,655,000, and $C$ from 4,800 to 7,600 lbs. per sq. in. Cubes of concrete made almost like that used for the Newark C.C.'s Rosendale cement gave $E_c$ about 538,000, while cubes of cement-mortar, and some of concrete, both made with National Portland cement, showed $E_c$ from 800,000 to 2,000,- - 600 lbs. per sq. in. The compressibility of brick piers 12 in. square in section and 16 in. high was also tested. They were made of common North River brick with mortar joints $\frac{1}{8}$ in. thick, and showed a value for $C'$ of about 1,555,000 lbs., while at elastic limit $C''$ was on the average 1,300 lbs. per sq. in. A diagram showing the relationship between the elastic modulus ($E_c$), compressibility ($C'$), and elastic limit ($C''$) for different materials. TORSION. 233 CHAPTER II. TORSION. 214. Angle of Torsion and of Helix. When a cylindrical beam or shaft is subjected to a twisting or torsional action, i.e. when it is the means of holding in equilibrium two couples in parallel planes and of equal and opposite moments, the longitudinal axis of symmetry remains straight and the elements along it experience no stress (whence it may be said that the line is "rigid"), while the lines originally parallel to it assume the form of helices, each element of which is distorted in its angles (originally right angles), the amount of distortion being assumed proportional to the radius of the helix. The directions of the Diagram showing a twisted shaft with helical elements. 213 faces of any element were originally as follows: two radial, two in consecutive transverse sections, and the other two tangent to two consecutive circular cylinders whose common axis is that of the shaft. E.g. in Fig. 212 we have an unstrained shaft, while in Fig. 213 it holds the two 234 MECHANICS OF ENGINEERING. couple (of equal moment $P_a = Q_b$) in equilibrium. These couples act in parallel planes perpendicular to the axis of the prism and a distance, $l$, apart. Assuming that the transverse section remains plane, the prism is in torsion, any two faces along $m$, which in fig. 212 was entirely right-angled, is now distorted. Two of its angles have been increased, two diminished, by an amount $\delta$, the angle between the helix and a line parallel to the axis. Supposing $m$ to be the most distant of any element from the axis, this distance being $c$, any other element at a distance $z$ from the axis experiences an angular distortion $-\frac{z}{c} \delta$. If now we draw $O B$ parallel to $OA$ the angle $B O B'$, ---a, is called the Angle of Torsion, while $\delta$ may be called the helix angle; the former lies in a transverse plane, the latter in a plane tangent to the cylinder. Now $$\tan \delta = (\text{linear arc } B' B') / l$$ but lin. arc $BB' = ca$; hence, putting $\delta$ for tan $\delta$, ($\delta$ being small) $$\delta = \frac{ca}{l} \quad \ldots \ldots \ldots (1)$$ ($\delta$ and $a$ both in $\pi$ measure). 215. Shearing Stress on the Elements. The angular distortion, or shearing strain, $\delta$, of any element (bounded as already described) is due to the shearing stresses exerted on it by its neighbors on the four faces perpendicular to the tangent plane of the cylindrical surface on which the element is situated. Consider these neighboring elements of an outside element removed, and the stresses put in; the latter are accountable for the distortion of the element and so A diagram showing a cross-section of a cylindrical prism with a helical groove. The diagram includes labels for various points and lines: \( P_a \), \( P_b \), \( m \), \( z \), \( c \), \( l \), \( \delta_1 \), \( \delta_2 \), \( \delta_3 \), \( \delta_4 \), and \( \delta \). The text "Fig. 215" is below the diagram." (1) TOESSION. 235 hold it in equilibrium. Fig. 214 shows this element "free." Within the elastic limit $p$ is known to be proportional to $p_{u}$ the shearing stress per unit of area on the faces whose relative angular positions have been changed. That is, from (1) § 208, $\delta = p_{u} E_{uu}$ whence, see (1) of § 214, $$p_{u} = \frac{e = E}{l} \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \\ In (2) $p_{u}$ and $e$ both refer to a surface element, $e$ being the radius of the cylinder, and $p$, the greatest intensity of shearing stress existing in the shaft. Elements lying nearer the axis suffer shearing stresses of least intensity in proportion to their radial distances, i.e., to their helix-angles. That is, the shearing stress on that face of the element which forms a part of a transverse section and whose distance from the axis is $x$, is $p_{u} = -\frac{e}{l} p_{u}$ per unit of area, and the total shear on the face is $pdF$, $dF$ being the area of the face. 216. Torsional Strength.—We are now ready to expose the full transverse section of a shaft under torsion, to deduce formulae of practical utility. Making a right section of the shaft of Fig. 213 anywhere between the two couples and considering the left hand portion as a free body, the forces holding it in equilibrium are the two forces $F$ of the left-hand couple plus an infinite number of shearing forces, each equal to its circle's radius times the cross section exposed by the removal of the right-hand portion. The cross section is assumed to remain plane during torsion, and is composed of an infinite number of $dF$'s, each being the area of an exposed face of an element; see Fig. 215. 236 MECHANICS OF ENGINEERING. A diagram showing a section of a shaft with a shear force applied to it. The diagram includes labels for various parts such as "P," "dF," "s," "Oo," "A," and "dF." The text below the diagram states: "Each elementary shearing force = dF, p.dF, and s is its lever arm about the axis Oo. For equilibrium, Z (mom.) about the axis Oo must = 0; i.e., in detail." or, reducing, $$\frac{P}{e} \int dF = Pa; \text{ or, } \frac{Pa}{e} = Pa \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ Eq. (3) relates to torsional strength, since it contains $p_e$, the greatest shearing stress induced by the torsional couple, whose moment $Pa$ is called the Moment of Torque, the stresses in the cross section forming a couple of equal and opposite moment. $f_t$ is recognized as the Polar Moment of Inertia of the cross section, discussed in § 94; $e$ is the radial distance of the outermost element, and $r$ is the radius for a circular shaft. 217. Torsional Stiffness.—In problems involving the angle of torsion, or deformation of the shaft, we need an equation connecting $Pa$ and $a$, which is obtained by substituting in eq. (3) the value of $p_e$ in eq. (2), whence $$\frac{aIE}{l} - Pa \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (4)$$ TOBSSION. 237 From this it appears that the angle of torsion, $a$, is proportional to the moment of torsion, $P_0$, within the elastic limit; a must be expressed in $\pi$-measures. Trautwine cises 1° (i.e. $a = 0.0174$) as a maximum allowable value for shafts. 218. Torsional Resilience is the work done in twisting a shaft from an unstrained state until the elastic limit is reached in the outermost elements. If in Fig. 218 we imagine the right-hand extremity to be fixed, while the other end is gradually twisted through an angle $a$, each force $P$ of the couple must be made to increase gradually from a zero value up to the value $P_0$, corresponding to $a$. In this motion each end of the arm describes a space $-\frac{1}{2}aa$, and the mean value of the force is $\frac{1}{2}P_0$ (compare § 196). Hence the work done in twisting is $$U_t = \frac{1}{2}P_0 \times \frac{1}{2}aa \times 2\pi = \frac{1}{2}P_0aa.$$ (5) By the aid of preceding equations, (5) can be written $$U_t = \frac{a^2EJ}{2}, \text{or } = \frac{P_0Vl}{2EI}, \text{or } = \frac{p_0II}{2EJ}.$$ (6) If for $p$, we write $S'$ (Modulus of safe shearing) we have for the safe resilience of the shaft $$U_t' = \frac{S'II}{2EJ}.$$ (7) If the torsional elasticity of an originally unstrained shaft is to be the means of arresting the motion of a moving mass whose weight is $G$ (large compared with the parts intervening) and velocity $=v$, we write (§ 133) $$U_t' = \frac{G}{g} \cdot \frac{\varphi^2}{2};$$ as the condition that the shaft shall not be injured. 238 MECHANICS OF ENGINEERING. **219. Polar Moment of Inertia.—For a shaft of circular cross-section (see § 94) $I = \frac{1}{2} \pi r^2$; for a hollow cylinder $I = \frac{1}{2} \pi (r_1^2 - r_2^2)$; while for a shaft of rectangular section $I = \frac{1}{6} b h^3$, $b$ being the side length and $h$, $bh$. For a cylindrical hollow, $\epsilon m r$, the greater radius. For a square, $\epsilon m \sqrt{\frac{2}{\pi}}$. **220. Non-Circular Shafts.—If the cross-section is not circular it becomes warped, in torsion, instead of remaining plane. Hence the foregoing theory does not strictly apply. The following results have been obtained by many, however, cover many of these cases. (See § 708 of Thompson and Tait's Natural Philosophy ; also, Prof. Burr's Elas- ticity and Strength of the Materials of Engineering). His results give for a square shaft (instead of the) $$\frac{d^2E_s}{dt^2} = Pa$$ of eq. (4) of § 217), $$Pa = 0.841 \frac{d^2E_s}{dt^2}$$ (1) and $$Pa = \frac{d^2E_s}{dt^2}$$ instead of eq. (9) of § 216, p., being the greatest shearing stress. The elements under greatest shearing strain are found at the middles of the sides, instead of at the corners, when the prism is of square or rectangular cross-section. The warping of the cross-section in such a case is easily verified by the student by twisting a bar of india-rubber in his fingers. **221. Transmission of Power.—Fig. 316. Suppose the cog-wheel $A$ to cause $A$, on the same shaft, to revolve uniformly and overcome a resistance $Q$, the pressure of the teeth of another cog-wheel, $B$ being driven by still another wheel. The shaft $AB$ is un-** A diagram showing two cog-wheels connected by a shaft. One wheel is labeled A and B, with arrows indicating rotation direction. The other wheel is labeled C and D. Pn. 316 TORISSON. 239 der torsion, the moment of torsion being $P_0 = Qb$. ($P$ and $Q$, the bearing reactions have no moment about the axis of the shaft). If the shaft makes $n$ revolutions per unit-time, the work transmitted (transmitted), not expended in twisting the shaft whose angle of torsion remains constant, corresponding to $P_0$ per unit-time, i.e. the Power, $$L = P \cdot 2\pi a \cdot u = 2\pi u P_0$$ (8) To reduce $L$ to Horse Power ($§ 192$), we divide by $N$, the number of units of work per unit-time constituting one H. P. in the system of units employed, i.e., $$H. P. = \frac{2\pi P_0}{N}$$ For example $N = 33,000$ ft.-lbs. per minute, or $= 396,000$ inch-lbs. per minute; or $= 550$ ft.-lbs. per second. Usually the rate of rotation of a shaft is given in revolutions per minute. But eq. (8) happens to contain $P_0$ the moment of torsion acting to maintain the constant value of the angle of torsion, and since for safety (see eq. § 216) $P_0 = SL + e$, with $S = \frac{1}{2}\pi r^2$ and $e = r$ for a solid circular shaft, we have for such a shaft (Safe), H. P. = $\frac{\pi^2 S r^2}{N}$ (9) which is the safe H. P., which the given shaft can transmit at the given speed. $S$ may be made 7,000 lbs. per sq. inch for wrought iron; 10,000 for steel, and 5,000 for cast-iron. If the value of $P_0$ fluctuates periodically, as when a shaft is driven by a connecting rod and crank, for (H. P.) we put $m \times (H. P.)$, $m$ being the ratio of the maximum to the mean torusional moment; $m$ is about $1/4$ under ordinary circumstances (Cotterill). A diagram showing a shaft with a connecting rod and crank. 240 MECHANICS OF ENGINEERING. 232. Autographic Testing Machine.—The principle of Prof. Thurston's invention bearing this name is shown in Fig A diagram showing the principle of an autographic testing machine. Fig. 232. 217. The test-piece is of a standard shape and size, its central cylinder being subjected to torsion. A jaw, carrying a handle (or gear-wheel turned by a worm) and a drum on which paper is wrapped, takes a firm hold of one end of the test-piece, while the other end is held by a jaw rigidly connected with a heavy pendulum carrying a pencil free to move axially. By a continuous slow motion of the handle the pendulum is gradually deviated more and more from the vertical, through the intervention of the test-piece, which is thus subjected to an increasing torsional moment. The axis of the test-piece lies in the axis of motion. This motion of the pendulum by means of a properly curved guide, $W$, causes an axial (i.e., parallel to axis of test-piece) deviation $\beta$ equal to that of the pendulum, as well as an angular deviation $\dot{\beta}$ equal to that of the pendulum, and this axial distance $CF_1 = T_1$ of the pencil from its initial position measures the moment of torsion $P_a = P_c \sin \beta$. As the piece twists, the drum and paper move relatively to the pencil through an angle $sUo$ equal to the angle TOBSSION. 241 of torsion $a$ so far attained. The abscissa $so$ and ordinate $sT$ of the curve thus marked on the paper, measure, when the paper is unrolled, the values of $a$ and $Pa$ through all the stages of the torsion. Fig. 218 shows typical A diagram showing the relationship between torque ($P_a$), angle ($\theta$), and radius ($r$) in a torsion problem. The x-axis represents torque, the y-axis represents angle, and the z-axis represents radius. The curve shows how torque changes with angle and radius. curves thus obtained. Many valuable indications are given by these strain diagrams as to homogeneity of composition, ductility, etc., etc. On relaxing the strain at any stage within the elastic limit, the pencil retracts its path; but if beyond that limit, a new path is taken called an "elasticity-line," in general parallel to the first part of the line, and showing the amount of angular recovery, $BC$, and the permanent angular set, $OB$. 223. Examples in Torsion.—The modulus of safe shearing strength, $S'$, as given in § 221, is expressed in pounds per square inch; hence these two units should be adopted throughout in any numerical examples where one of the above values for $S'$ is used. The same statement applies to the modulus of shearing elasticity, $E$, in the table of § 210. EXAMPLE 1.—Fig. 216. With $P = 1$ ton, $a = 3$ ft., $l = 10$ ft., and the radius of the cylindrical shaft $r = 2.5$ inches, required the max. shearing stress per sq. inch, $p_s$, the shaft being of wrought iron. From eq. (3) § 216 $$p_s = \frac{Pae}{I} = \frac{2000 \times 36 \times 2.5}{(1/4\pi \times (2.5)^3)} = 930 \text{ lbs. per sq. inch},$$ which is a safe value for any ferrous metal. A diagram showing the relationship between torque ($P_a$), angle ($\theta$), and radius ($r$) in a torsion problem. The x-axis represents torque, the y-axis represents angle, and the z-axis represents radius. The curve shows how torque changes with angle and radius. 242 MECHANICS OF ENGINEERING. **EXAMPLE 2.—What H. P. is the shaft in Ex. 1 transmitting, if it makes 50 revolutions per minute? Let $e$ = number of revolutions per unit of time, and $N$ = the num- ber of units of work per unit of time constituting one horse-power. Then H. P. = $P_{\text{horse}} \times e \times N$, which for the foot-pound-minute system of units gives H. P. = $2,000 \times 60 \times 2 \times 3 + 33,000 = 57 \frac{1}{2}$ H. P. **EXAMPLE 3.—What different radius should be given to the shafts in Ex. 1, if two radii at its extremities, originally parallel, are to make an angle of $2^\circ$ when the given moment of torsion is acting, the strains in the shaft remaining constant. From eq. (4) § 217, and the table 210, with $e = \frac{\pi}{180} = 0.035$ radians (i.e. $\pi$-measure), and $r_1 = r_2$, we have $$\frac{2,000 \times 36 \times 120}{\frac{\pi}{180} \times 0.035 \times 9,000,000} = 17.45 . . . r = 2.04$ inches. (This would bring about a different $p$, but still safe.) The foregoing is an example in efficiency. **EXAMPLE 4.—A working shaft of steel (solid) is to transmit 4,000 H. P. and make 60 rev. per minute, the maximum twisting moment being $1\frac{1}{2}$ times the average; required its diameter. $$d = 14.74$$ inches. Ans. **EXAMPLE 5.—In example 1, $p_1 = 2,930$ lbs. per square inch; what tensile stress does this imply on a plane at $45^\circ$ with the pair of planes on which $p_1$ acts? Fig. 219 shows A diagram showing a cross-section of a solid shaft with a rectangular cross-section ABCD and two inclined planes ABE and CDF making angles of $45^\circ$ with the horizontal plane. Fig. 219 A diagram showing a cross-section of a solid shaft with a rectangular cross-section ABCD and two inclined planes ABE and CDF making angles of $45^\circ$ with the horizontal plane. Fig. 220 TORISSION. 243 a small cube, of edge $=dx$, (taken from the outer helix of Fig. 215.) free and in equilibrium, the plane of the paper being tangent to the cylinder; while 220 shows the portion $BDC$, also free, with the unknown total tensile stress $pdz^2/dy$ acting on the newly exposed rectangle of area $=dx \times dx \times dy$, p being the unknown stress per unit of area. From symmetry the stress on this diagonal plane has no shearing component. Putting 2 components normal to $BD$ = 0, we have $$pdz^2dy = 2dFp_{\text{cos}45^\circ} \cdot adz \cdot dp_{\text{sin}45^\circ} \cdot p = p_{\text{c}} \cdot . \quad (1)$$ That is, a normal tensile stress exists in the diagonal plane $BD$ of the cubic element equal in intensity to the shearing stress on one of the faces, i.e., = 2,930 lbs. per sq. in. in this case. Similarly in the plane $AC$ will be found a compressive stress of 2,930 lbs. per sq. in. If a plane surface had been exposed making any other angle than 45° with the face of the cube, the shearing would still be zero, but the tangential and normal stresses each less than $p_c$ per sq. inch. Hence the interior dotted cube in 219, if shown "free" is in tension in one direction, in compression in the other, and with no shear, these normal stresses having equal intensities. Since $S'$ is usually less than $T'$ or $C'$, if $p_c$ is made = $S'$, the tensile and compressive actions are not injurious. It follows therefore that when a cylinder is in torsion any helix at an angle of $\alpha$ with its axis is a line of tensile, or of compressive stress, according as it is a right or left handed helix, or vice versa. **Example 6.**—A solid and a hollow cylindrical shaft, of equal length, contain the same amount of the same kind of metal, the solid one fitting the hollow of the other. Compare their torsional strengths, used separately. The solid shaft has only $\frac{1}{2}$ the strength of the hollow one. Ans. A diagram showing a small cube with edges dx and dy, and a diagonal plane BD. The cube is free from shearing stress on one face. 244 MECHANICS OF ENGINEERING. CHAPTER III FLEXURE OF HOMOGENEOUS PRISMS UNDER PERPENDICULAR FORCES IN ONE PLANE. 224. Assumptions of the Common Theory of Flexure.—When a prism is bent, under the action of external forces perpendicularly to it and in the same plane with each other, it may be assumed that the longitudinal fibres are in tension on the convex side, in compression on the concave side, and that the amount of this contraction by the elements is proportional to their distances from a plane intermediate between, with the understanding that the flexure is slight and that the elastic limit is not passed in any element. This "common theory" is sufficiently exact for ordinary engineering purposes if the constants employed are properly determined by a wide range of experiments, and involves only assumptions of practical nature as possible, consistently with practical facts. These assumptions are as follows, (for prisms, and for solids with variable cross sections, when the cross sections are similarly situated as regards a central straight axis) and are approximately borne out by experiment: (1) The external or "applied" forces are all perpendicular to the axis of the piece and lie in one plane, which may be called the force-plane; the force-plane contains the axis of the piece and cuts each cross-section symmetrically; (2) The cross-sections remain plane surfaces during flexure; (3) There is a surface (or, rather, sheet of elements) which is parallel to the axis and perpendicular to the force-plane, and along which the elements of the solid-ex- FLEXURE 245 perience no tension nor compression in an axial direction, this being called the Neutral Surface. (4.) The projection of the neutral surface upon the force plane (or a plane) being called the Neutral Line or Elastic Curve, the bending or flexure of the piece is so slight that an elementary division, $ds$, of the neutral line may be put $=dx$, its projection on a line parallel to the direction of the axis before flexure; (5.) The elements of the body contained between any two consecutive sections, whose intersections with the neutral surface are the respective Neutral Axes of the sections, experience elongations (or contractions, accord- ing as they are situated on one side or the other of the neutral surface), in an axial direction, whose amounts are proportional to their distances from the neutral axis, and indicate corresponding tensile or compressive stresses; (6.) $E=E_0$; (7.) The dimensions of the cross-section are small com- pared with the length of the piece; (8.) There is no shear perpendicular to the force plane on internal surfaces perpendicular to that plane. In the locality where any one of the external forces is applied, local stresses are of course induced which demand separate treatment. These are not considered at present. 225. Illustration.—Consider the case of flexure shown in Fig. 221. The external forces are three (neglecting the Fig. 221. FORCE PLANE 246 MECHANICS OF ENGINEERING. weight of the beam), viz.: $P_a$, $P_b$ and $P_c$. $P_a$ and $P_b$ are loads, $P_c$ the reaction of the support. The force plane is vertical. $N_L$ is the neutral line or elastic curve. $N_A$ is the neutral axis of the cross-section at $m$; this cross-section, originally perpendicular to the sides of the prism, is during flexure $\perp$ to their tangent planes. The section $QNR$ is, on the other hand, the side view $QNR$ of any cross-section is perpendicular to the neutral line. In considering the whole prism free we have the system $P_a$, $P_b$ and $P_c$ in equilibrium, whence from $ΣY=0$ we have $P_a + P_b + P_c$ and from $Σ(mom.\ about\ O)=0$, $P_d = P_e$. Hence given $P_c$ we may determine the other two external forces. A reaction such as $P_c$ is sometimes called a supporting force. The elements above the neutral surface $N_LOZS$ are in tension; those below in compression (in an axial direction). 226. The Elastic Forces.—Conceive the beam in Fig. 221 separated into two parts by any transverse section such as $QA$, and the portion $N_1ON$, considered as a free body in Fig. 222. Of this free body the surface $QAB$ is one of A diagram showing a beam with a transverse section cut through it, labeled QAB. The diagram includes various lines and points labeled with letters and numbers. Page 221 FLEXURE. 247 the bounding surfaces, but was originally an internal surface of the beam in Fig. 221. Hence in Fig. 223 we must put in the stresses acting on all the $dF$ elements of area of $QdL$. These stresses represent the actions of the body taken away upon the body which is left, and according to assumptions (5), (6) and (8) consist of normal stresses (tension or compression) proportional per unit of area, to the distance, $x$, of the $dF$'s from the neutral axis, and of shearing stress, perpendicular to the force-plane (which in most cases will be vertical). The intensity of this shearing stress on any $dF$ varies with the position of the $dF$ with respect to the neutral axis, but the law of its variation will be investigated later ($\S$ 233 and 254). These stresses, called the Elastic Forces of the cross-section exposed, and the external forces $P_{n}$ form a system in equilibrium. We may therefore apply any of the conditions of equilibrium proved in § 33. 227. The Neutral Axis Contains the Centre of Gravity of the Cross-Section.—Fig. 223. Let $e$ = the distance of the outermost element of the cross-section from the neutral axis, and the normal stress per unit of area upon it be $\sigma e$, whether tension or compression. Then by assumptions (5) and (6), $\S$ 224, the intensity of normal stress on any $dF$ is $\frac{e}{c} \cdot P$ and the actual normal stress on any $dF$ is $\frac{e}{c} \cdot pdF$. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) This equation is true for $dF$'s having negative $x$'s, i.e. on the other side of the neutral axis, the negative value of the force indicating normal stress of the opposite character, for if the relative elongation (or contraction) of two adjacent strips at that point be equal, then below, the neutral surface, the stresses producing the changes in length are also the same, provided $E_1=E_2$; see §§ 184 and 201. 248 **MECHANICS OF ENGINEERING.** For this free body in equilibrium put $XX=0$ ($X$ is a horizontal axis). Put the normal stresses equal to their $X$ components, the flexure being so slight, and the $X$ component of the shears = 0 for the same reason. This gives (see eq. (1) ) $$\int_{e}^{L} pdF = 0 \text{ ; i.e. } P \int dFx = 0 \text{ ; or, } P \cdot Fx = 0 \quad (2)$$ In which $i =$ distance of the centre of gravity of the cross-section from the neutral axis, from which, though unknown in position, the $x$'s have been measured (see eq. (4) § 23). In eq. (2) neither $p-e$ nor $F$ can be zero : $i$ must = 0; i.e. the neutral axis contains the centre of gravity. Q. E. D. [If the external forces were not all perpendicular to the beam this result would not be obtained, necessarily.] 228. The Shear.—The “total shear,” or simply the “shear,” in the cross-section is the sum of the vertical shearing stresses on the respective $dF$’s. Call this sum $J$, and we shall have from the free body in Fig. 222, by putting $XY=0$ ($Y$ being vertical) $$P - P_x - J = 0; J = P_x - P_x \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ That is, the shear equals the algebraic sum of the external forces acting on one side (only) of the section considered. This result implies nothing concerning its mode of distribution over the section. 229. The Moment.—By the “Moment of Flexure” or simply the “Moment,” at any cross-section is meant the sum of the moments of the elastic forces of the section, taking the neutral axis as an axis of moment. In this summation these forces are taken as a shear alone, the shear taking no part, having no lever arm about the neutral axis. Hence, Fig. 222, the moment of flexure, A diagram showing a free body in equilibrium with a horizontal axis labeled X, and a vertical axis labeled Y. The diagram includes various lines and points representing different forces and moments. FLEXURE. 249 $$f\left(\frac{z}{e}pdF\right)=\frac{p}{e}f^{\prime}dFd=\frac{pf}{e}\ .\ \ .\ \ (4)$$ This function, $f'dFd$, of the cross-section or plane figure is the quantity called Moment of Inertia of a plane figure, § 85. For the free body in Fig. 222, by putting $Y(mom.s)$ about the neutral axis $M(A)=0$, we have then $$\frac{pf}{e}=P_{x_1}+P_{x_2}=0,\ \mathrm{or}\ \mathrm{in}\ \mathrm{general},\ \frac{pf}{e}=M\ .\ (5)$$ in which $M$ signifies the sum of moments, about the neutral axis of the section, of all the forces acting on the free body considered, exclusive of the elastic forces of the exposed section itself. 230. Strength in Flexure.—Eq. (5) is available for solving problems involving the strength of beams and girders, since it contains $p$, the greatest normal stress per unit of area to be found in the section. In the preceding present chapter, where all the external forces are perpendicular to the prism or beam, and have therefore no components parallel to the beam, i.e. to the axis $X$, it is evident that the normal stresses at any section, as $QB$ Fig. 222, are equivalent to a couple; for the condition $\Sigma X=0$ falls entirely upon them and cannot be true unless the resultant of the tensions is equal, parallel, and opposite to that of the compressions. These two equal and parallel resultants, not being in the same line, form a couple ($\Sigma Y$), which we may call the stress-couple. The moment due to this couple is evidently zero; hence $\Sigma Y=0$, it is further evident that the remaining forces in Fig. 222, viz.: the shear $J$ and the external forces $P_x$ and $P_y$ are equivalent to a couple of equal and opposite moment to the one formed by the normal stresses. A diagram showing a beam with forces applied at different points along its length. 250 MECHANICS OF ENGINEERING. 231. Flexural Stiffness.—The neutral line, or elastic curve, containing the centres of gravity of all the sections, was originally straight; its radius of curvature at any point, as $N$, Fig. 223, during flexure may be introduced as follows. Draw two consecutive $ds$ sections, originally parallel, but not intersecting so that the intersection $C$, found by prolonging them sufficiently, is the centre of curvature of the $ds$ (pnt = $dx$) which separates them at $N$, and $CO = \rho$ the radius of curvature of the elastic curve at $N$. From the similar triangles $U'UG$ and $GNC$ we have $\frac{d}{dx} : dx :: \rho$, in which $\frac{d}{dx}$ is the elongation, $UU'$ of a portion, originally equal to $dx$, of the outer fibre. But the relative elongation $= \frac{d}{dx}$ of the latter is, by §184, within the elastic limit, $\frac{E}{F} = \frac{P}{B} = \frac{e}{f}$ and eq. (5) becomes $$\frac{EI}{F} = M$$ (6) From (6) the radius of curvature can be computed. $E$ = value of $E_1 = E_2$ as ascertained from experiments in bending. To obtain a differential equation of the elastic curve, (6) may be transformed thus, Fig. 223. The curve being very flat, consider two consecutive $ds$ with equal $dx$’s; they may be put = their $dx$’s. Produce the first to intersect the $dy$ of the second, thus cutting off the $dy'$, i.e. the difference between two consecutive $dy$’s. Drawing a perpendicular to each $ds$ at its left extremity, the centre of curvature $O$ will be found by inter- section, and thus the radius of curvature $\rho$. The two shaded triangles have their small angles A diagram showing two shaded triangles with small angles at their vertices. Fig. 223 FLEXURE. 251 equal, and $dy$ is nearly perpendicular to the prolonged $ds$ ; hence, considering them similar, we have $$p : dx : ds : dy : \frac{1}{p} \cdot \frac{dy}{ds}$$ and hence from eq. (6) we have (approx.) $\pm EI \frac{dy}{ds} = M$ (7) as a differential equation of the elastic curve. From this the equation of the elastic curve may be found, the deflections at different points computed, and an idea thus formed of the stiffness. All beams in the present chapter being *prismatic* and *homogeneous* both $E$ and $I$ are the same (i.e. constant) at all points of the elastic curve. In using (7), it must be remembered that the length $ds$ of the beam before flexure, which must be slight; the minus sign in (7) provides for the case when $dy + ds$ is essentially negative. 233a. Resilience of Flexure—If the external forces are made to increase gradually from zero up to certain maximum values, some of them may do work by reason of their points of application moving through certain distances due to the yielding, or flexure, of the body. If at the beginning and also at the end of this operation the body is at rest, this work has been expended on the elastic resistance of the body, and an equal amount, called the work of resilience (or springing-back), will be restored by the elasticity of the body, if released from the external forces, provided the elastic limit has not been passed. The energy thus temporarily stored is of the potential kind; see §§ 148, 180, 196 and 218. 233b. Distinction Between Simple, and Continuous, Beams (or "Girders").—The external forces acting on a beam consist 252 MECHANICS OF ENGINEERING. generally of the loads and the "reactions" of the supports. If the beam is horizontal and rests on two supports only, the reactions of those supports are easily found by elementary statics (§ 36) alone, without calling into account the fact that it is a continuous beam, to be a Simple Beam, or girder; whereas if it is in contact with more than two supports, being "continuous," therefore, over some of them, it is a Continuous Girder (§ 271). The remainder of this chapter will deal only with simple beams. ELASTIC CURVES. 233. Case I. Horizontal Prismatic Beam, Supported at Both Ends. With a Central Load, Weight of Beam Neglected—Fig. 224. First considering the whole beam free, we find each A diagram showing a horizontal prismatic beam supported at both ends with a central load. The beam is labeled AOB, and the reaction to be is labeled M. The line AB is the neutral line. Fn. 224. $4$ cm. reaction to be = $\frac{1}{2}P$. AOB is the neutral line; required the equation of the portion OB referred to O as origin, and to the tangent line through any section m on the axis of X. To do this consider the section Bm as any section m on the right of O and the near support, in Fig. 225. The forces holding this free body in equilibrium A diagram showing a horizontal prismatic beam supported at both ends with a central load. The beam is labeled AOB, and the reaction to be is labeled M. The line AB is the neutral line. Fn. 225. A diagram showing a horizontal prismatic beam supported at both ends with a central load. The beam is labeled AOB, and the reaction to be is labeled M. The line AB is the neutral line. Fn. 226. ELASTIC CURVES. 253 are the one external force $\frac{1}{2}P$, and the elastic forces acting on the exposed surface. The latter consist of $J$, the shear, and the tensions and compressions represented in the figure by the equivalent "stress-couple". Selecting $W$, the vertical axis, $m_0$, $m_1$, and $m_2$ moments (that may not appear in the moment equation) and putting $Z(\mathrm{mom})=0$ we have $$\frac{F(l)}{\frac{1}{2}-x}-EI\frac{dy}{dx}=0 \cdot EI\frac{dy}{dx}=\frac{P}{2}\left(\frac{l}{2}-x\right) \cdot (1)$$ Fig. 226 shows the elastic curve $OB$ in its purely geometrical aspect, much exaggerated. For axes origin as in figure $dy+dx^2$ is positive. Eq. (1) gives the second $x$-derivative of $y$ equal to a function of $x$. Hence the first $x$-derivative of $y$ will be equal to the anti-derivative of that function, plus a constant, $C$. (By anti-derivative is meant the converse of derivative, sometimes called integral though not in the sense of summation). Hence from (1) we have ($EI$ being a constant factor remaining undisturbed) $$EI\frac{dy}{dx}=\frac{P}{2}\left(\frac{l}{2}-\frac{x^2}{2}\right)+C \cdot \quad (2)'$$ (3') is an equation between two variables $dy+dx$ and $x$, and holds good for any point between $O$ and $B$. $dy+dx$ denotes the tangent of $\alpha$, the slope, or angle between the tangent line and $X$. At $O$ the slope is zero, and $x$ also zero; hence at $O$ (3') becomes $$EI\times0=0+C$$ which enables us to determine the constant $C$, whose value must be the same at $O$ as for all points of the curve. Hence $C=0$ and (3') becomes 254 MECHANICS OF ENGINEERING. $$EI\frac{dy}{dx} = \frac{P}{2}\left(\frac{l}{2}-\frac{x^2}{2}\right) \cdot \cdot \cdot (3)$$ from which the slope, $\tan a$, (or simply $a$, in $x$-measure; since the angle is small) may be found at any point. Thus at $B$ we have $x=\frac{l}{2}$ and $dy+dx=a$, and $\therefore a = \frac{1}{16}\cdot\frac{PP}{EI}$ Again, taking the $x$-anti-derivative of both members of eq. (2) we have $$EIy = \frac{P}{2}\left(\frac{l^2}{4}-\frac{x^2}{6}\right)+C' \cdot \cdot \cdot (3')$$ and since at $O$ both $x$ and $y$ are zero, $C'$ is zero. Hence the equation of the elastic curve $OB$ is $$EIy = \frac{P}{2}\left(\frac{l^2}{4}-\frac{x^2}{6}\right) \cdot \cdot \cdot (3)$$ To compute the deflection of $O$ from the right line joining $A$ and $B$ in Fig. 224, i.e. $BK=d$, we put $x=\frac{l}{4}$ in $(3)$, $y$ being then $=d$, and obtain $$BK = d\cdot\frac{1}{48}\cdot\frac{PP}{EI} \cdot \cdot \cdot (4)$$ Eq. (3) does not admit of negative values for $x$; for if the free body of Fig. 225 extended to the left of $O$, the external forces acting would be $P$, downward, at $O$; and $\frac{l}{4}P$, upward, at $B$, instead of the latter alone; thus altering the form of eq. (1). From symmetry, however, we know that the curve $AO$, Fig. 224, is symmetrical with $OB$ about the vertical through $O$. ELASTIC CURVES. 235 233a. Load Suddenly Applied.—Eq. (4) gives the deflection $d$ corresponding to the force or pressure $P$ applied at the middle of the beam, and is seen to be proportional to it. If a load $G$ is suddenly applied at the middle of the beam, $P=G$; but if the load $G$, being initially placed at rest upon the unbent beam, is suddenly released from the external constraint necessary to hold it there, it sinks and deflects the beam, the pressure $P$ actually felt by the beam varying with the deflection as the load sinks. What is the ultimate deflection $d_{u}$? Let $P_{m}=$ the pressure between the load and the beam at the instant of maximum deflection. The work so far done in bending the beam $= \frac{1}{2} P_{m} d_{u}$. The potential energy given up by the load $= G d_{u}$ while the initial and final kinetic energies are both nothing. \therefore G d_{u} = \frac{1}{2} P_{m} d_{u} \quad \dots \quad (8) That is, $P_{m}=2G$. Since at this instant the load is subjected to an upward force of $2G$ and to a downward force of only $G$ (gravity) it immediately begins an upward motion, reaching the point whence the motion began, and thus the oscillation continues. We here suppose the elasticity of the beam unimpaired. This is called the "sudden" application of a load, and produces, as shown above, double the pressure on the beam which it does when gradually applied, and a double deflection. The work done by the beam in raising the weight again is called its "resilience." Similarly, if the weight $G$ is allowed to fall on the middle of the beam from a height $h$, we shall have $$G \times (k + d_u), \text{ or approx., } Gh = \frac{1}{2} P_m d_u;$$ and hence, since (4) gives $d_u$ in terms of $P_m$ $$G h = \frac{1}{96} \cdot \frac{P^2 m}{E l}; \text{ or } G h = \frac{24 E I r^2}{r}. \quad (6)$$ 256 MECHANICS OF ENGINEERING. This theory supposes the mass of the beam small compared with the falling weight. 234. Case II. Horizontal Prismatic Beam, Supported at Both Ends, Bearing a Single Eccentric Load. Weight of Beam Neglected.—Fig. 287. The reactions of the points of support, $P_1$ and $P_2$, are easily found by considering the whole beam free, and putting $F = P(l - x)$, $x$ being the distance $F_1 = F(l + l_t)$, and that $x(mom.) = 0$, whence $P_1 = P(l - l_t) + l_t$, $P_2$ and $P_1$ will now be treated as known quantities. The elastic curves $OC$ and $CB$, though having a common tangent line at $C$ (and hence the same slope $\alpha_c$), and a common ordinate at $C$, have separate equations and are both referred to the same origin and axes, as shown in the figure. The slope at $O_a$, and that at $B_a$, are unknown constants, to be determined in the progress of the work. Equation of OC—Considering as free a portion of the beam extending from $B$ to a section made anywhere on $OC$, $x$ and $y$ being the co-ordinates of the neutral axis of that section, we conceive the elastic forces put in on the exposed surface, as in the preceding problem, and put $\lambda(mom., about neutral axis of the section) = 0$ which gives (remembering that here $dy/dz$ is negative) $$E\frac{d^2y}{dx^2} = P(x - e) - P(l - x); \quad . . . \quad (1)$$ whence, by taking the $x$ anti-derivatives of both members $$hI\frac{dy}{dx} = P(x - \frac{x^2}{2}) - P(l - x - \frac{x^2}{2}) + C$$ To find $C$, write out this equation for the point $O$, where $dy/dx = \alpha_c$ and $x = 0$, and we have $C = EIta_c$; hence the equation for slope is FLEXURE ELASTIC CURVES. 257 $$EI\frac{dy}{dx}=P(lz-\frac{x^2}{2})-P_1(lz-\frac{x^2}{2})+EIa_0 \quad . \quad (2)$$ Again taking the $x$ anti-derivatives, we have from (3) $$Ef_y=P(\frac{lz-x^2}{2}-\frac{x^2}{6})-P_1(\frac{lz-x^2}{2}-\frac{x^2}{6})+EIa_xx+C'=0) \quad (3)$$ (at $O$ both $x$ and $y$ are =0, $C'=0$). In equations (1), (3), and (3) no value of $x$ is to be used <0 or >l, since for points in $CB$ different relations apply, thus Equation of CR—Fig. 227. Let the free body extend from $B$ to a section made anywhere on $OB$. $\Sigma$(moms.), as before, =0, gives $$EI\frac{dy}{dx}=-P_1(lz-x) \quad . \quad . \quad (4)$$ (N.B. In (4), as in (1), $EIdy+dx^2$ is written equal to a negative quantity because itself essentially negative; for the curve is concave to the axis $X$ in the first quadrant of the co-ordinate axes.) From (4) we have in the ordinary way (z-anti-deriv.) $$EI\frac{dy}{dx}=-P_1(lz-\frac{x^2}{2})+C'' \quad . \quad . \quad (5)$$ To determine $C''$, consider that the curves $CB$ and $OC$ have the same slope ($dy/dx$) at $Q$ where $x=zl$; hence put $z=l$ in the right-hand members of (3) and of (5) and equate the results. This gives $C''=\frac{1}{2}Pr+EIa_0$ and : $$EI\frac{dy}{dx}=\frac{Pr}{2}+EIa_0-P_1(lz-\frac{x^2}{2}) \quad . \quad (5)$$ and : $EIy=\frac{Pr}{2}x+Ela_0-P_1(\frac{lz-x^2}{2})+C'' \quad . \quad (6)$ 258 **MECHANICS OF ENGINEERING.** At $C$, where $x=l$, both curves have the same ordinate; hence, by putting $x=l$ in the right members of (3) and (6), and equating results, we obtain $C''=\frac{1}{2}PP$. .. (6) be comes $$Ely = \frac{1}{2}Pp + Eta - P\left(\frac{l^2-x^2}{2}\right)\frac{PP}{6}$$ (6) as the Equation of CR, Fig. 227. But $a_0$ is still an unknown constat, to find which write out (6) for the point $B$ where $x=l$, and $y=0$, whence we obtain $$a_0 = \frac{1}{6EI_0}(PP-3Pl+2Pt)$$ (7) $a_0$ = a similar form, putting $F_1$ for $P_0$, and $(l-l)$ for $l$. **235. Maximum Deflection in Case II—Fig. 227.** The ordinate $y_o$ of the lowest point is thus found. Assuming $l>\frac{1}{2}l_0$, it will occur in the curve $OC$. Hence put the $dy/dx$ of that curve, as expressed in equation (2), =0. Also for $a_0$ write its value from (7), having put $P_1=PP+l_0$, and we have $$P(lz-\frac{l^2}{2}) - P\frac{l}{l_0}(l-z-\frac{x^2}{2}) + \frac{1}{2}\frac{Pl}{l_0}(l-3l_0+2lt)=0$$ whence [for max. y] = $\sqrt{\frac{1}{2}}(2l-l)$ Now substitute this value of $x$ in (3), also $a_0$ from (7), and put $F_1=PP+l_0$, whence $$Max.\ Deflec.=ymax=\frac{1}{6}\cdot\frac{P}{EI_0}(P-3pl+2lt)\cdot\sqrt{\frac{1}{2}}(2l-l).$$ **236. Case III. Horizontal Prismatic Beam Supported at Both Ends and Bearing a Uniformly Distributed Load along its Whole Length.—(The weight of the beam itself, if considered,) FLEXURE. ELASTIC CURVES. 259 constitutes a load of this nature.) Let $l$ = the length of the beam and $w$ = the weight, per unit of length, of the loading; then the load coming upon any point $x$ will be $=wx$, and the whole load $=wl$. By hypothesis $w$ is constant. Fig. 228. From symmetry we know that the A diagram showing a beam with a load applied at various points along its length. The diagram includes labels for different sections of the beam and the load. Fig. 228. reactions at $A$ and $B$ are each $\frac{1}{2}wl$, that the middle $O$ of the neutral line is its lowest point, and the tangent line at $O$ is horizontal. Conceiving a section made at any point $m$ of the neutral line at a distance $x$ from $O$, consider as free the portion of beam on the right of $m$. The forces holding this portion in equilibrium are $\frac{1}{2}wl$, that acting at $B$, the elastic force pressing on exposed surface at $m$, viz., the tensional pressure, $\frac{1}{2}wl$, and a couple $J$, the total shear; and a portion of the load, $wx(\frac{1}{2}l-x)$. The sum of the moments of these latter forces about the neutral axis of $m$, is the same as that of their resultant; (i.e., their sum, since they are parallel), and this resultant acts in the middle of the length $\frac{1}{2}l-x$. Hence the sum of these moments $=wx(\frac{1}{2}l-x)(\frac{1}{2}l-x)$. Now putting $E$ (mom. about neutral axis of $m$)=0 for this free body, we have $$Et\frac{dy}{dx}=\frac{1}{2}wx(\frac{1}{2}l-x)-\frac{1}{2}wx(\frac{1}{2}l-x)^{2};$$ i.e., $Et\frac{dy}{dx}=\frac{1}{2}wx(\frac{1}{2}l-x)$ . . . . . . . . . . . . . . . . . . . . . . . . . (1) 259 260 MECHANICS OF ENGINEERING. Taking the $x$-anti-derivative of both sides of (1), $$EI \frac{dy}{dx} = \frac{1}{2} w (\frac{d^2y}{dx^2}) + (C=0) \quad (2)$$ as the equation of slope. (The constant is $=0$ since at $O$ both $dy/dx$ and $z$ are $=0$.) From (2), $$EI y = \frac{w}{2} (\frac{1}{2} P x^2 - \frac{1}{3} M x^3) + [C'=0] \quad . . . \quad (3)$$ which is the equation of the elastic curve; throughout, i.e., it admits any value of $x$ from $-\frac{L}{2}$ to $\frac{L}{2}$. This is an equation of the fourth degree, one degree higher than those for the Curves of Cases I and II, where there were no distributed loads. If $w$ were not constant, but proportional to the ordinates of an inclined right line, eq. (3) would be of the fifth degree; if $w$ were proportional to the ordinate of a straight line parallel to the axis vertical, (3) would be of the sixth degree; and so on. By putting $x=\frac{L}{2}$ in (3) we have the deflection of $O$ below the horizontal through $A$ and $B$, viz.: (with $W=$ total load = $wL$) $$d = \frac{5}{384} \frac{w^2 L^4}{EI} = \frac{5}{384} \frac{WP L^4}{EI} \quad . . . \quad (4)$$ 237. Case IV. Castilever—A horizontal beam whose only support consists in one end being built in a wall, as in Fig. 229(a), or supported as in Fig. 229(b) is sometimes called a cantilever. Let the student prove that in Fig. 229(a) with a single end load $P$, the deflection of $B$ below the tangent at $O$, $\frac{1}{3} P L^4 / EI$, is such that statement applies to Fig. 229(b); but the tangent at $O$ is not horizontal if the beam was originally so. It can also be proved that the slope at $B$, Fig. 229(a) (from the tangent at $O$) is A diagram showing a cantilever beam with a single end load \( P \). The deflection of point \( B \) below the tangent at point \( O \) is shown. Fig. 229(a) FLEXURE ELASTIC CURVES 261 $$a = \frac{PF}{2ET}$$ The greatest deflection of the elastic curve from the right line joining $AB$, in Fig. 239(a), is evidently given by the equation for $y$ max. in § 255, by writing, instead of $P$ of that equation, the reaction at $O$ in Fig. 239(b). This assumes that the max. deflection occurs between $A$ and $O$. If it occurs between $O$ and $B$ put $(l_1-l)$ for $l$. If in Fig. 239(a) the loading is uniformly distributed along the beam at the rate of $w$ pounds per linear unit, the student may also prove that the deflection of $B$ below the tangent at $O$ is $$d = \frac{1}{2} w l^2 + EI = \frac{1}{6} WP$$ $$EI$$ 238. Case V. Horizontal Prismatic Beam Bearing Equal Terminal Loads and Supported Symmetrically at Two Points—Fig. 239. Weight of beam neglected. In the preceding cases we have made use of the approximate form $EIr_y + dLz$ in determining the forms of elastic curves. In the present case the elastic curve from $O$ to $C$ is more directly dealt with by employing the more exact expression $EI/r + p$ (see § 231) for the moment of the stress-couple in any section. The reactions at $O$ and $C$ are each $-p$, from symmetry. Considering free a portion of the beam extending from $A$ to any section $m$ between $O$ and $C$ (Fig. 239) we have, by putting $\Sigma$ (mom. about neutral axis of $m$) = 0, $$P(l+z) - \frac{EI}{r} - Pz = 0 : p = \frac{EI}{Pl}$$ 262 MECHANICS OF ENGINEERING. That is, the radius of curvature is the same at all points of $OC$; in other words $OC$ is the arc of a circle with the above radius. The upward deflection of $F$ from the right line joining $O$ and $C$, $\Delta F$, can be estimated from knowledge of this fact. This is left to the student as also the value of the slope of the tangent line at $O$ (and $C$). The deflection of $D$ from the tangent at $C = \frac{1}{2}PP + EI$, as in Fig. 229(a). SAFE LOADS IN FLEXURE. 239. Maximum Moment.—As we examine the different sections of a given beam under a given loading we find different values of $p$, the normal stress per unit of area in the outer element, as obtained from eq. (5) § 229, viz: $$\frac{pl}{e} = M.$$ (1) in which $I$ is the "Moment of Inertia" (§ 85) of the plane figure formed by the section, about its neutral axis, $e$ the distance to the most distant (or outer) fibre from the neutral axis, and $M$ the moment of inertia about the neutral axis, of all the forces acting on the free body of which the section in question is one end, exclusive of the stresses on the exposed surface of that section. In other words $M$ is the sum of the moments of the forces which balance the stresses of the section, these moments being taken about the neutral axis of the section under examination. For the prismatic beams of this chapter $e$ and $I$ are the same at all sections, hence $p$ varies with $M$ and becomes a maximum when $e = \frac{1}{2}PP$. This point is called the location of the "dangerous section," or section of maximum $M$, and the amount of that maximum value may be determined by inspection and trial, this being the only method (except by graphic) if the external forces are detached. FLEXURE SAFE LOADS. 263 If, however, the loading is continuous according to a definite algebraic law the calculus may often be applied, taking care to treat separately each portion of the beam between two consecutive reactions of supports, or detached loads. As a graphical representation of the values of $M$ along the beam in any given case, these values may be conceived laid off as vertical ordinates (according to some definite scale, e.g., centimetres) on the vertical axis of moment (see the inch of paper) from a horizontal axis just below the beam. If the upper fibres are in compression in any portion of the beam, so that that portion is convex downwards, these ordinates will be laid off below the axis, and vice versa; for it is evident that at a section where $M = 0$, it also $-0$, i.e., the character of the normal stress in the outermost fibre changes (from tension to compression, or vice versa) when $M$ changes sign. It is also evident from eq. (6) § 251 that the radius of curvature changes sign, and consequently the curvature changes sign. The ordinates of the moment diagram form a Moment Diagram, and the extremities a Moment Curve. The maximum moment, $M_{\text{max}}$, being found, in terms of the loads and reactions, we must make the p of the "dangerous section," where $M - M_{\text{max}}$ equal to a safe value $R'$, and thus may write $$\frac{RI}{c} = M_{\text{max}} \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad \ldots $$ Eq. (2) is available for finding any one unknown quantity, whether it be a load, span, or some one dimension of the beam, and is concerned only with the strength and not with the stiffness of the beam. If it is satisfied in any given case, the normal stress on all elements in all sections is known to be = or $< R'$, and the design is therefore safe in that one respect. As to danger arising from the shearing stresses in any 204 MECHANICS OF ENGINEERING. section, the consideration of the latter will be taken up in a subsequent chapter and will be found to be necessary only in beams composed of a thin web uniting two flanges. The total shear, however, denoted by $J$, bears to the moment $M$, an important relation of great service in determining $M_w$. This relation, therefore, is presented in the next article. 240. The Shear is the First x-Derivative of the Moment—Fig. 233. ($x$ is the distance of any section, measured parallel to the beam from an arbitrary origin). Consider as free a vertical slice of the beam included between any two consecutive vertical sections whose distance apart is $dx$. The forces acting on this slice form a couple of the two internal surfaces now laid bare, and, possibly, a portion, $wdx$, of the loading, which at this part of the beam has some intensity $-w$ lbs per running linear unit. Putting $\Sigma$ (mom. about axis $N'$) = 0 we have (noting that since the tensions and compressions of section $N$ form a couple, the sum of their moments about $N'$ is just the same as about $N$.) $$\frac{pI}{e} + p' \frac{I'}{e} + Jdx + w dx \frac{dx}{2} = 0$$ But $p' = M$, the Moment of the left hand section, $\frac{p'I}{e} = M$, that of the right; whence we may write, after dividing through by $dx$ and transposing, $$\frac{M - M'}{dx} = J + w \frac{dx}{2} \quad \text{i.e.} \quad \frac{dM}{dx} = J; \quad (3)$$ for $\frac{dM'}{dx}$ vanishes when added to the finite $J$, and $M - M' = dM$ increment of the moment corresponding to the increment, $dx$, of $x$. This proves the theorem. A diagram showing a cross-section of a beam with various labeled parts. **FLEXURE.SAFE LOADS.** 265 Now the value of $x$ which renders $M$ a maximum or minimum would be obtained by putting the derivative $dM/dx=0$; hence we may state as a **Corollary.** At sections where the moment is a maximum or minimum the shear is zero. The shear $J$ at any section is easily determined by considering from the portion of beam from the section to either end of the beam the ordinates (vertical components) = 0. In this article the words maximum and minimum are used in the same sense as in calculus; i.e., graphically, they are the ordinates of the moment curve at points where the tangent line is horizontal. If the moment curve be reduced to a straight line, or a series of straight lines, it has no maximum or minimum in the strict sense just stated; nevertheless the relation is still practically borne out by the fact that at the points of greatest and least ordinates on the diagram the tangent line does sign suddenly. This is best shown by drawing a shear diagram, whose ordinates are laid off vertically from a horizontal axis and under the respective sections of the beam. They will be laid off upward or downward according as $J$ is found to be upward or downward, when the free body considered extends from the section towards the right. In these diagrams the moment ordinates are set off on an arbitrary scale of so many inch-pounds, or foot-pounds, to the inch, or some other unit of force, on a scale of so many pounds to the inch of paper. The scale on which the beam is drawn is so many feet, or inches, to the inch of paper. **241. Safe Load at the Middle of a Prismatic Beam Supported at the Ends—Fig. 234.** The reaction at each support is $\frac{1}{4}P$. Make a section $n$ at any distance $\frac{1}{4}$ from $B$. Consider the portion $AB$ free, putting in the proper elastic and external forces. The weight of beam is neglected. From $I(mom.\ about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .about\ .$ we have 266 MECHANICS OF ENGINEERING. $$\frac{P}{e} = \frac{P}{2x}; \text{i.e., } M = \frac{1}{2}Px$$ Evidently $M$ is proportional to $x$, and the ordinates representing it will therefore be limited by the straight line A diagram showing a triangle O'RA with a horizontal line through its midpoint. From symmetry, another triangle O'RA forms the other half of the moment diagram. From inspection, the maximum $M$ is seen to be in the middle where $x = \frac{1}{2}L$, and hence $$M_{\text{max}} = M_0 = \frac{1}{2}PL \quad . . . . (1)$$ Again by putting $Z(\text{vert. comp.)} = 0$, for the free body $nB$ we have $$J = \frac{P}{2}$$ and must point downward since $\frac{P}{e}$ points upward. Hence the shear is constant and $\frac{1}{2}P$ at any section in the right hand half. If $nB$ be taken in the left half we would have, $nB$ being free, from $Z(\text{vert. com.)} = 0$, $$J - P = \frac{1}{2}P = \frac{1}{2}I$$ FLEXURE, SAFE LOADS. 267 the same numerical value as before; but $J$ must point up-ward, since $\frac{c}{a}$ at $B$ and $J$ at $A$ must balance the downward $P$ at $A$. At $A$, then, the shear changes sign suddenly, that is, passes through the value zero; also at $A$, $M$ is a maximum, thus illustrating the statement in § 240. Notice the shear diagram in Fig. 234. To find the safe load in this case we write the maximum value of the normal stress, $pE$, a safe value (see table in a suitable article) and solve the equation for $P$. But the maximum value of $p$ is in the outer fibre at $A$, since $M$ for that section is a maximum. Hence $$\frac{RI}{e} = \frac{1}{4}PL$$ (2) is the equation for safe loading in this case, so far as the normal stresses in any section are concerned. EXAMPLE.—If the beam is of wood and has a rectangular section with width $b=2$ in., height $h=4$ in., while its length $l=10$ ft., required the safe load, if the greatest normal stress is limited to 1,000 lbs. per sq. in. Use the pound and inch. From § 90 $I=\frac{bh^3}{12}=\frac{1}{12}\times2\times64=10.68$ biquad. inches, while $e=2$ in. $$\therefore P = \frac{4ET}{le} = \frac{4\times1000\times10.68}{120\times2} = 177.7 \text{ lbs.}$$ 242. Safe Load Uniformly Distributed along a Prismatic Beam Supported at the Ends—Let the load per linear unit of the length of beam be $w$ (this can be made to include the weight of the beam itself). Fig. 235. From symmetry, each reaction = $\frac{1}{2}wl$. For the free body no we have, putting $\Sigma(mom.$ about $n)$ = 0, $$\frac{pl}{e} - \frac{wl}{2}(x-w) + M = \frac{wl}{2}(x-x')$$ 268 MECHANICS OF ENGINEERING. which gives $M$ for any section by making $x$ vary from 0 to $l$. Notice that in this case the law of loading is continuous along the whole length, and that hence the moment curve is continuous for the whole length. Diagram showing a beam with a load at its center. Fig. 3a To find the shear $J$, at $n$, we may either put $F$ (vert. comp.) = 0 for the free body, whence $J = \frac{1}{2}w(l - x)$, and must therefore be downward for a small value of $x$; or, employing § 24), we may write out $dM/dx$, which gives $$\frac{dM}{dx} = \frac{w}{2}(l - 2x)$$ (1) the same as before. To find the max. $M_n$ or $M_o$, put $J = 0$, which gives $x = \frac{l}{2}$. This indicates a maximum, for when substituted in $dM/dx$, i.e., in $-w$, a negative result is obtained. Hence $M_o$ occurs at the middle of the beam and its value is $$M_o = \frac{1}{2}wl^2 ; . . . \frac{RI}{e} = \frac{1}{2}wl^2$$ (2) the equation of safe loading. $W$ = total load = wl. It can easily be shown that the moment curve is a por- FLEXURE, SAFE LOADS. 269 tion of a parabola, whose vertex is at $A''$ under the mid- dle of the beam, and axis vertical. The shear diagram consists of ordinates to a single straight line inclined to its axis and crossing it, i.e., giving a zero shear, under the middle of the beam. If a frictionless dove-tail joint with vertical faces were introduced at any locality in the beam and thus divided the beam into two parts, the presence of $J$ would be made manifest by the downward slipping of the left hand part on the right hand part if the joint were on the right of the middle, and vice versa if it were on the left of the middle. This shows why the ordinates in the two halves of the shear diagram have opposite signs. The greatest shear is close to either support and is $d = \frac{1}{2}wt$. 243. Primatic Beam Supported at Its Extremities and Loaded in Any Manner. Equation for Safe Loading—Fig. 236. Given the loads $P_0$, $P_1$, and $P_2$ whose distances from the right sup- port are $x_0$, $x_1$, and $x_2$, find the equation for safe loading; i.e., find $M_x$ and write it = $RI - e$. If the moment curve were continuous, i.e., if $M$ were a continuous function of $x$ from end to end of the beam, we could easily find $M_x$ by making $a = x$. But since we cannot sub- stitute the resulting value of $x$ in the expression for $M$. But in the present case of detached loads, $J$ is not zero, necessarily, at any section of the beam. Still there is some one section where it changes sign, i.e., passes sud- denly through the value zero, and this will be the section of greatest moment (though not a maximum in the strict sense used in calculus). By considering any portion $s - t$ as free, $J$ is found equal to the Reaction at 0 diminished by the Loads Occurring Between s and t. The reaction at B is A diagram showing a beam with loads P0, P1, and P2 applied at different distances from the right support. The diagram also shows a moment curve M(x) that intersects the x-axis at some point between x0 and x1. 236 270 **MECHANICS OF ENGINEERING.** $$P_s = (P_A + P_D + P_L) + l$$ obtained by treating the whole beam as free (in which case no elastic forces come into play) and putting 2Tmom. about Oy=0; while that at O is $$F_1 = F_2 + F_3 - F_4$$ If $s$ is taken anywhere between $O$ and $A$, we have $$E = F, J = P_s - P_t$$ $$" " " " " F = H, J = P_s - P_t$$ $$" " " " " H = B, J = P_s - P_t - P_t$$ This last value of $J$ also is the reaction at the other support, $B$. Accordingly, the shear diagram is seen to consist of a number of horizontal steps. The relation $J = dM/dz$ is such that the slope of the moment curve is proportional to the ordinate of the shear diagram, and that for each load, when the slope of the moment curve there is a sudden change in the shear ordinate. Hence in the present instance, $J$ being constant between any two consecutive loads, the moment curve reduces to a straight line between the same loads, this line having a different inclination under each of the portions into which the beam is divided by the loads. Under each load the slope of the moment curve and the ordinate of the shear diagram change suddenly. In Fig. 250 the shear passes through the value zero, i.e., $P_s = P_t$, at $l_1$, and then changes to $-P_s$, supposing to find that $P_s > P_t$, while $P_s - P_t > P_t$ is in the present case. Considering $FO$, then, as free, we find $M_a$ to be $$M_a = P_A - P_d(l_1 - l_1) \text{ and the equation for safe loading is}$$ $$\frac{RI}{c} = P_d - P_d(l_1 - l_1) \quad (1)$$ (i.e., if the max $M$ is at F). It is also evident that the greatest shear is equal to the reaction at one or the other support, whichever is the greater, and that the moment at either support is zero. The student should not confuse the moment curve, which FLEXURE, SAFE LOADS. 271 is entirely imaginary, with the neutral line (or elastic curve) of the beam itself. The greatest moment is not necessarily at the section of maximum deflection of the neutral line (elastic curve). For the case in Fig. 236 we may therefore state that the max. moment, and consequently the greatest tension or compression in the outer fibre, will be found in the section under that load for which the sum of the loads (including this load itself) between it and either support first equals or exceeds the reaction of that support. The amount of this moment is then obtained by treating as free either of the two portions of the beam into which this section divides the beam. 244. Numerical Example of the Preceding Article.—Fig. 237. Given $P_1$, $P_2$, $P_3$ equal to $\frac{1}{2}$ ton, 1 ton, and 4 tons, re- Figure 237. Fig. 237. respectively; $l_1$=5 feet, $l_2$=7 feet, and $l_3$=10 feet; while the total length is 15 feet. The beam is of timber, of rectangular cross-section, the horizontal width being $b$=10 inches, and the value of $T$ (greatest safe normal stress) = $\frac{1}{2}$ ton per sq. inch, or 1,000 lbs. per sq. inch. 272 272 MECHANICS OF ENGINEERING. Requirea the proper depth $h$ tor the beam, for safe load- ing. Solution.—Adopting a definite system of units, viz., the inch-ton-second system, we must reduce all distances such as $s$, etc., to inches, express all forces in tons, write $R = \frac{1}{4}$ (tons per sq. inch), and interpret all results by the same system. Moments will be in inch-tons, and shears in tons. [N.B. In problems involving the strength of materials the inch-ton-second system is used throughout this book, since any stress expressed in lbs., or tons, per sq. inch, is numerically 144 times as small as if referred to the square foot.] Making the whole beam free, we have from moma. about O, $P_0 = 1\left(\frac{1}{4} \times 60 + 1 \times 84 + 4 \times 120\right) = 3.3$ tons.; $P_1 = 5.5 - 3.3 = 2.2$ tons. The shear anywhere between $O$ and $E$ is $J = P_0 = 2.2$ tons. " " " " " " $E$ and $F$ is $J = 2.2 - \frac{1}{4} = 1.7$ tons. The shear anywhere between $F$ and $H$ is $J = 2.2 - \frac{1}{4} - 1 = 0.7$ tons. The shear anywhere between $H$ and $B$ is $J = 2.2 - \frac{1}{4} - 1 - \frac{1}{4} = 3.3$ tons. Since the shears changes sign on passing $H$, : the max. moment is at $H$, whence taking HO free, we have $$M_{at H} = M_0 = \frac{1}{4} \times 120 - \frac{1}{4} \times 60 - 1 \times 30 = 198$$ inch-tona. For safety $M_n$ must $\geqslant \frac{RT}{e}$, in which $R = \frac{1}{4}$ ton per sq. inch, $e = y_2 h = \frac{y_2}{h}$ of unknown depth of beam, and $T_{80} = \frac{5}{8} B^2$, with $B = 10$ inches $\therefore \frac{y_2}{h} \times 10^4 = 198$; or $h^2 = 237.6$.:. $h = 15.4$ inches. 245. Comparative Strength of Rectangular Beams.—For such a beam, under a given loading, the equation for safe load- ing is $$\frac{HI}{e} = M_n$$ i.e. $\frac{y_2}{h} H b h^2 = M_n$$... (1) FLEXURE SAFE LOADS. 273 whence the following is evident, (since for the same length, mode of support, and distribution of load, $M_m$ is proportional to the safe loading.) For rectangular prismatic beams of the same length, same material, same mode of support and same arrangement of load : (1) The safe load is proportional to the width of beams having the same depth ($h$). (2) The safe load is proportional to the square of the depth of beams having the same width ($b$). (3) The safe load is proportional to the depth of beams having the same volume (i. e., the same $S_b$). (4) It is understood that the sides of the section are horizontal and vertical respectively and that the material is homogeneous.) 246. Comparative Stiffness of Rectangular Beams—Taking the deflection under the same loading as an inverse measure of the stiffness, and noting that in §§ 233, 235, and 236, this deflection is inversely proportional to $I = \frac{1}{3} b h^3 =$ the "moment of inertia" of the section about its neutral axis, we have thus: For rectangular prismatic beams of the same length, same material, same mode of support, and same loading : (1) The stiffness is proportional to the width for beams of the same depth. (2) The stiffness is proportional to the cube of the height for beams of the same width ($b$). (3) The stiffness is proportional to the square of the depth for beams of equal volume ($b h^3$). (4) If the length alone vary, the stiffness is inversely proportional to the cube of the length. 247. Table of Moments of Inertia—These are here recapitulated for the simpler cases, and also the values of $e$, the distance of the outermost fibre from the axis. Since the stiffness varies as $I$ (other things being equal), 274 MECHANICS OF ENGINEERING. while the strength varies as $I + e$, it is evident that a square beam has the same stiffness in any position (899), while its strength is greatest with one side horizontal, for then $e$ is smallest, being $\frac{1}{2}b$. Since for any cross-section $I = \int dF^{'z}$, in which $z$=the distance of any element, $dF$, of area from the neutral axis, a beam is made both stiffer and stronger by throwing most of its material into two flanges united by a vertical web, thus making a box-girder (Fig. 230). But not without limit, for the web must be thick enough to cause the flanges to act together as a solid of continuous substance, and, if too high, is liable to buckle sideways, thus requiring lateral stiffening. These points will be treated later. A diagram showing a box girder. **SECTION.**
I e
Rectangle, width = b, depth = A (vertically) $\frac{1}{16}ba^3$ $\frac{1}{3}A$
Rectangle, symmetrical about neutral axis. See Fig. 230. $\frac{1}{16}a_1^3A_1 - \frac{1}{3}A_2$ $\frac{1}{3}A_1$
Triangle, width = b, height = h, neutral axis parallel to base of triangle. $\frac{1}{3}ba^2h$ $\frac{1}{3}A$
Circle of radius r Ker$^2r$ r
Ring of concentric circles. Fig. 230.d) Ker$(r_1^2-r_2^2)$ r1
Rhombus. Fig. 230.c) A diagonal which is vertical. $\frac{1}{4}ba^2$ $\frac{1}{4}A$
Square with side b vertical. $\frac{1}{4}b^3$ b3
- = - at 45° with horiz. $\frac{1}{4}b^3$ b3
248. Moment of Inertia of I-beams, Box-girders, Etc.—In common with other large companies, the N. J. Steel and FLEXURE. SAFE LOADS. 275 Iron Co. of Trenton, N. J. (Cooper, Hewitt & Co.) manu- facture prismatic rolled beams of wrought-iron variously called I-beams, deck-beams, rails, and "shape iron," (in- cluding channels, angles, tees, etc., according to the form of section.) See fig. 239 for these forms. The company A diagram showing different shapes of beams. Fig. 239. publishes a pocket-book giving tables of quantities rela- ting to the strength and stiffness of beams, such as the safe loads for various spans, moments of inertia of their sections in various positions, etc., etc. The moments of inertia of I-beams and deck-beams are computed accord- ing to §§ 92 and 93, with the latter as linear unit. The I-beams range from about 6 to 30 inches deep, the deck- beams being about 7 and 8 in. deep. For beams of still greater stiffness and strength com- binations of plates, channels, angles, etc., are riveted to- gether, forming "built-beams," or "plate girder." The proper design for the riveting of such beams will be ex- amined later. For the present the parts are assumed to act together as a continuous mass. For example, Fig. 240 shows a "box-girder," formed by two "channels" and two plates riveted together. In this case the neutral axis is to be horizontal it becomes the neu- tral axis. Let \( C \) = the moment of iner- tia of one channel (as given in the pocket-book mentioned) about the axis \( N \) perpendicular to the web of the chan- nel. Then the total moment of inertia of the combination is (nearly) \[ I_x = 2C + 2hd^2 - 4d^2(d - \frac{h}{2})^2 \] (1) A diagram showing a box-girder made up of two channels and two plates. 276 MECHANICS OF ENGINEERING. In (1), $b$, $t$, and $d$ are the distances given in Fig. 240 ($d$ extends to the middle of plate) while $d'$ and $t'$ are the length and width of a rivet, the former from head to head (i.e., $d$ and $t'$ are the dimensions of a rivet-hole). For example, in a box-girder bridge the channel is formed of two 15-inch channels and two plates 10 inches wide and 1 inch thick, the rivet holes $\frac{3}{4}$ in. wide and $1\frac{3}{4}$ in. long. That is, $b=10$; $t=1$; $d=\frac{3}{4}$; $t'=\frac{3}{4}$; and $d'=1\frac{3}{4}$ inches. Also from the pocket-book we find that for the channel in question, $C=376$ biquadratic inches. Hence, eq. (1) $$I_4 = 752 + 2 \times 10 \times 1 \times 64 - 4 \times \left(\frac{3}{4} \times \frac{3}{4} \times (8-5)\right) = 1737 \text{ biquadr.in.}$$ Also, since in this instance $e = \frac{5}{8}$ inches, and 19000 lbs. per sq. inch (or 6 tons per sq. in.) is the value for $R$ (=greatest safe normal stress on the outer element of any cross-section) used by the Trenton Co. (for wrought iron), we have $$\frac{Rf}{e} = \frac{19000 \times 1737}{8.5} = 2451700 \text{ inch-lbs.}$$ That is, the box-girder can safely bear a maximum moment, $M_{max} = 2451700$ inch-lbs. = 1225.8 inch-tons, as far as the normal stresses in any section are concerned. (Proper provision for the shearing stresses in the section, and in the rivets, will be considered later). 249. Strength of Cantilevers.—In Fig. 241 with a single concentrated load $P$ at the projecting extremity, we easily find the moment at $s$ to be $M = Pz$, and the max. moment to occur at its value being $\frac{Pz^2}{2}$. The shear is constant, and $P$ at all sections. The moment and shear diagrams are drawn in accordance with these results. Fig. 241 Fig. 242 FLEXURE. SAFE LOADS. 277 If the load $W = wL$ is uniformly distributed on the cantilever, as in fig. 243, by making $nO$ free we have, putting $E(mom. about. O) = 0$, $$\frac{pI}{e} = \frac{x^2}{y} - M_{max} = \frac{wL}{y} = \frac{wL}{y}.$$ Hence the moment curve is a parabola, whose vertex is at $O'$ and axis vertical. Putting $x' (vert. comp.) = 0$ we obtain $J = wL$. Hence the shear diagram is a triangle, and the max. $J' = wL = W$. 250. Résumé of the Four Simple Cases.—The following table shows the values of the deflections under an arbitrary load $P$, or $W$, (within elastic limit), and of the safe load;
Cantilevers. Beams with two end supports.
With one end fixed. With end load at $x_1$.
Fig. 241 Fig. 242
Deflection $\frac{PL^3}{3EI}$ $\frac{PL^3}{3EI}$
Safe load (from $P'$) $\frac{PL}{2}$ $\frac{PL}{2}$
Max. wall $\frac{PL}{8}$ $\frac{PL}{8}$
Relative strength 1 1
Relative stiffness (under same load) 1 1
Relative strength (under safe load) 1 1
Max. shear = Max. (and) bending. $P$, int wall $W$, ext wall
$M_P$, int suply. $M_W$, ext supply.
also the relative strength, the relative stiffness (under the same load), and the relative stiffness under the safe load, for the same beam. The max. shear will be used to determine the proper web-thickness for I-beams and "built-girders." The student should carefully fill in the foregoing table, noting especially the relative strengths, stiffnesses, and stiffnesses under safe load, of the same beam. Thus, a beam with two end supports will bear a double 278 MECHANICS OF ENGINEERING. load, if uniformly distributed instead of concentrated in the middle, but will deflect $\frac{1}{4}$ more whereas with a given load uniformly distributed the deflection would be only $\frac{1}{4}$ of that caused by the same load in the middle, provided the elastic limit is not surpassed in either case. 251. B; etc. For Various Materials—The formula $P_d = M_{\text{max}} \cdot e$ from which in any given case of flexure we can compute the value of $p_o$, the greatest normal stress in any outer element, provided all the other quantities are known, holds good theoretically within the elastic limit only. Still, some experimenters have used this formula for the rupture of beams by flexure, calling the value of $p_o$ thus obtained the Modulus of Rupture, $R$. $R$ may be found to differ considerably from both the $T$ or $C$ of § 208 with some materials and forms, being frequently much larger. This might be expected, since even supposing the relative extension or compression (i.e., strain) of the fibres to be proportional to the distance from the neutral axis, as the load increases toward rupture, the corresponding stresses, not being proportional to these strains beyond the elastic limit, no longer vary directly as the distances from the neutral axis; and the neutral axis does not pass through the centre of gravity of the section, necessarily. The following table gives average values for $R$, $R'$, $R''$, and $E$ for the ordinary materials of construction, $E$, the modulus of elasticity for use in the formula for deflection, is given by experiments in flexure, and is nearly the same as $E_1$ and $E_2$. In any example involving $R$, $e$ is usually written equal to the distance of the outer fibre from the neutral axis, whether that fibre is to be in tension or compression; since in most materials not only is the tensile equal to the compressive stress for a given strain (relative extension or contraction) but the elastic limit is reached at about the same strain both in tension and compression. $$P_d = M_{\text{max}} \cdot e$$ FLEXURE. SAFE LOADS. 279 **TABLE FOR USE IN EXAMPLES IN FLEXURE.**
Timber Cast Iron Wrought Iron Steel
M.A.S. safe stress in outer S. 4,000 lbs. in comp. 32,000 lbs. 32,000 lbs. 32,000 lbs.
bow - Pyth. per sq. inch. 1,000 18,500 lbs. 18,500 lbs. 18,500 lbs.
Stress in outer fibre at E.M. 17,600 lbs. 17,600 lbs.
Hunt - Pyth. per sq. inch. 35,000 lbs. 35,000 lbs. and upward.
" Modul. of Resilience " 4,000 40,000 90,000 125,000
- R.-Bra. per sq. inch. 12 25,000
K - Modul. of Elasticity, 1,000,000 17,500,000 32,000,000 32,500,000
-Hla. per sq. inch L.M.G.L.
In the case of cast iron, however, (see § 29) the elastic limit is reached in tension with a stress of $9,999$ lbs. per sq. inch and a relative extension of $\frac{a}{b}$ of one per cent., while in compression the stress must be about double to reach the elastic limit, the relative change of form (strain) being also double. Hence with cast iron beams, once largely used but now almost entirely displaced by rolled wrought iron beams, the neutral axis was displaced by making the outer fibre on the compressed side twice as far from the neutral axis as that on the stretched side. Thus, Fig. 243, cross-sections with unequal flanges were used, so proportioned that the centre of gravity was twice as near to the outer fibre in tension as to that in compression, i.e., $e_1 = e_2$, in other words more material was placed in tension than in compression. The third figure shows (elastic limit), that at $A$, since it is twice as far from the neutral axis and on the other side, is contracted twice as much as $A$ is extended; i.e., is under a compressive strain double the tensile strain at $A$, but in accordance with the above figures its state of stress is proportionally as much within the elastic limit as that of $A$. Steel beams are gradually coming into use, and may ultimately replace those of wrought iron. *It is interesting to note that $K'$ ranged from 25,566 to 38,666 and the elastic limit was reached with less stress in the larger than in the smaller beams. Also, for the same beam $K'$ decreased with larger spans.* 280 MECHANICS OF ENGINEERING. The great range of values of $R$ for timber is due not only to the fact that the various kinds of wood differ widely in strength, while the behavior of specimens of any one kind depends somewhat on age, seasoning, etc., but also to the circumstance that the size of the beam under experiment has much to do with the result. The experiments of Prof. Lanza at the Mass. Institute of Technology in 1881 were made on full size lumber (spruce), of dimensions 3 by 6 inches, and were used in buildings, and gave much smaller values of $R$ (from 3,300 to 8,700 lbs. per sq. inch) than had previously been obtained. The loading employed was in most cases a concentrated load midway between the two supports. These low values are probably due to the fact that in large specimens of ordinary lumber the continuity of its substance is more or less broken by cracks, knots, etc., the higher values of most other experimenters having been obtained with well-staggered, selected planks, from one foot to six feet in length. The value $R = 12,000$ lbs. per sq. inch is employed by the N. J. Iron and Steel Co. in computing the safe loads for their rolled wrought iron beams, with the stipulation that the beams (which are high and of narrow width) must be secure against yielding sideways. If such is not the case the ratio of the actual safe load to that computed with $R = 12,000$ takes the form $p_{w} / p_{e}$ as the span increases. The lateral security referred to may be insured by the brick arch-filling of a fire-proof floor, or by light lateral bracing with the other beams. 252 Numerical Examples—Example I.—A square bar of wrought iron, 1⅛ in. in thickness is bent into a circular arc whose radius is 200 ft., the plane of bending being parallel to the side of the square. Required the greatest normal stress $p_w$ in any outer fiber. Solution. From §§ 230 and 231 we may write $$\frac{EI}{a} = \frac{pI}{c} \therefore p = E \cdot \frac{E}{c} \cdot p_e$$ *For "neir soft steel beams this company uses 16,000 lbs. per sq. inch.* A diagram showing a square bar bent into a circular arc. FLEXURE SAFE LOADS. 281 For the unit inch and pound (viz. those of the table in § 251) we have $e = \frac{1}{4}$ in., $\rho = 2,400$ lb., and $E = 25,000,000$ lbs. per sq. inch, and 1: $$p = p_{n} = \frac{1}{4} \times 25,000,000 + 2,400 \times 7,812 \text{ lbs. per sq. in.},$$ which is quite safe. At a distance of $\frac{1}{4}$ inch from the neutral axis, the normal stress is $= (\frac{1}{4} \times \frac{1}{4}) p_{n} = \frac{1}{16} p_{n} = 5,208$ lbs. per sq. in. (If the force-plane (i.e., plane of bending) were parallel to the diagonal of the square, $e$ would be $\sqrt{2}$ times greater than $\frac{1}{4}$.) The maximum stress $p_{m} = (7,812 \times \sqrt{2})$ lbs. per sq. in.; § 238 shows an instance where a portion, $OC$, Fig. 231, is bent in a circular arc. **Example 2.** A hollow cylindrical cast-iron pipe of radii $3/4$ and 4 inches * is supported at its ends and loaded in middle (see Fig. 234). Required the safe load, neglecting the weight of the pipe. From the table in § 250 we have for safety $$P = 4 \frac{I r}{l e}$$ From § 251 we put $R = 6,000$ lbs. per sq. in.; and from § 247 $I = \frac{\pi}{6}(r_1^4 - r_2^4)$; and with these values, $r_1$ being $=\frac{1}{4}$, $r_2 = 4$, $e = r_1 - r_2 = \frac{3}{4}$, and $l = 144$ inches (the inch must be the unit of length since $R = 6,000$ lbs. per sq. inch) we have $$P = 4 \times 6,000 \times \frac{3}{4} \cdot \frac{\pi}{6}(\frac{256 - 150}{\pi}) \cdot (144 \times 4) = P = 3,470 \text{ lbs.}$$ The weight of the beam itself is $G = F_T (\delta T)$, i.e., $$G = (r_1^4 - r_2^4) l e = \frac{\pi}{6}(16 - 13)(\frac{3}{4}) \cdot (144 \times \frac{3}{4}) = 443 \text{ lbs.}$$ (Notice that $r$, here, must be lbs., per cubic inch). This weight being a uniformly distributed load is equivalent to half as much, 221 lbs., applied in the middle, as far as the strength of the beam is concerned (see § 250). $P$ must be taken $= 3,249$ lbs. when the weight of the beam is considered. *And 12 feet in length between supports.* 282 MECHANICS OF ENGINEERING. **EXAMPLE 3.**—A wrought-iron rolled I-beam supported at the ends is to be loaded uniformly Fig. 285, the span being equal to 20 feet. Its cross-section, Fig. 244, has a depth of 16 inches, the height of 15 inches, a flange width of 5 inches. In the pocket book of the Trenton Co. it is called a 15-inch **light** I-beam, weighing 150 lbs. per yard, with a moment of inertia $=533$ bi-quad. inches about a gravity axis perpendicular to the web (i.e., when the web is vertical, the strongest position) and $=15$ biq. in. about a gravity axis parallel to the web (i.e., when the web is placed horizontally). First placing the web vertically, we have from § 250, $$W_1 = \text{Safe load}, \quad = 8 \frac{RI}{l_1}.$$ With $R = 12,000,$ $$I = 533, \quad l = 240 \text{ inches}, \quad e_1 = 7\frac{1}{2} \text{ inches}, \quad \text{this gives}$$ $$W_1 = [8 \times 12,000 \times 533] / [240 \times 7\frac{1}{2}] = 27,893 \text{ lbs}.$$ But this includes the weight of the beam, $G = 20$ ft. $\times$ lbs. $= 1,000$ lbs.; hence a distributed load of $26,908$ lbs., or $13.45$ tons may be placed on the beam (secured against lateral yielding by the pocket-book referred to gives $13.27$ tons as the safe load). Secondly, placing the web horizontal, $$W_2 = 8 \frac{RI}{l_2} = [8 \times 12,000 \times 15] / [240 \times 7\frac{1}{2}] = \frac{45}{823} \text{ of } W_1,$$ or only about $\frac{1}{12}$ of $W_1$. **EXAMPLE 4.**—Required the deflection in the first case of Ex. 3. From § 250 the deflection at middle is $$d_1 = \frac{5}{584} \cdot \frac{W_1 P}{EI_1} = \frac{5}{584} \cdot \frac{8 RI}{E l_1} = \frac{5}{584} \cdot \frac{P}{E} \cdot \frac{l_1}{e_1}.$$ A diagram showing a cross-section of an I-beam with dimensions: depth of 16 inches, height of 15 inches, flange width of 5 inches. Fig. 244. FLEXURE SAFE LOADS. 283 i.e., $d = \frac{5}{45} \cdot \frac{19,000}{95,000,000} \cdot \frac{(240)^2}{\pi^2}$ (inch and pound) :. $d = 0.384$ in. **EXAMPLE 5.—A rectangular beam of yellow pine, of width $b$=4 inches, is 20 ft. long, rests on two end supports, and is to carry a load of 1,200 lbs. at the middle; required the proper depth $h$. From § 250 $$P = \frac{4RI}{le} = \frac{4 \cdot 6P \cdot h^2}{l^2} \cdot \frac{1}{\frac{1}{h}}$$ :. $h = 6P/(4Rl)$. For variety, use the inch and ton. For this system of units $P$=0.60 tons, $R$=0.50 tons per sq. in., $l$=240 inches and $b$=4 inches. :. $h = (6 \times 0.6 \times 240)/(4 \times 0.5 \times 4)$=108 sq. in. i.e. $h$=10.4 in. **EXAMPLE 6.—Suppose the depth in Ex. 5 to be determined by the condition that the deflection shall be $\frac{1}{12}$th of the span or length. We should then have from § 250 $$d = \frac{300}{600} l = \frac{Pt}{EI}$$ Using the inch and ton, with $E$=1,200,000 lbs. per sq. in., which = 600 tons per sq. inch, and $I$=$y_{ab}^2$, we have $$P = \frac{500 \times 0.6 \times 240 \times 240 \times 12}{48 \times 600 \times 4} = 1875.$$ i.e. $h$=12.2 in. As this is > 10.4 the load would be safe, as well. **EXAMPLE 7.—Required the length of a wren's iron pipe supported at its extremities, its internal radius being $2\%$ in., the external radius being $3\%$ in., the reduction in cross-section due to weight may equal $\frac{1}{12}$th of the length. $$L = 579.6$$ in. Ams. **EXAMPLE 8.—Fig. 245. The wall is 6 feet high and one foot thick, of common brick work (see § 7) and is to be borne by an I-beam in whose outer fibres no greater normal stress than 8,000 lbs. per sq. inch is allowable. If a number of I-beams is available, A diagram showing a section of a wall with a beam spanning across it. Pa. 36. 284 MECHANICS OF ENGINEERING. ranging in height from 6 in. to 15 in. (by whole inches), which one shall be chosen in the present instance, if their cross-sections are similar figures, the moment of inertia of the 15-inch beam being 800 biquad. inches? The 12-inch beam. Ans. SHEARING STRESSES IN FLEXURE. 853. Shearing Stresses in Surfaces Parallel to the Neutral Surface.--If a pile of boards (see Fig. 246) is used to support a load, the boards being free to slip on each other, it is noticeable that the ends overlap, although the boards A diagram showing two piles of boards with overlapping ends. Fig. 246. A diagram showing two piles of boards with overlapping ends. Fig. 247. are of equal length (now see Fig. 247); i.e., slipping has occurred along the surfaces of contact, the combination being no stronger than the same board side by side. If, however, they are glued together, piled as in the former figure, the slipping is prevented and the deflection is much less under the same load $F$. That is, the compound shearing stress is smaller than when a pile of loose boards, but the tendency to slip still exists and is known as the "shearing stress in surfaces parallel to the neutral surface." Its intensity per unit of area will now be determined by the usual "free-body" method. In Fig. 248 let $AN'$ be a portion, considered free, on the left of any A diagram showing a free-body diagram for determining shearing stress. Fig. 248. SHEAR IN FLEXURE. 285 section $N'$, of a prismatic beam slightly bent under forces in one plane and perpendicular to the beam. The moment equation, about the neutral axis at $N$, gives $$\frac{pI}{c} = M' \text{; whence } p' = \frac{Me}{I} \text{ . . . . . (1)}$$ Similarly, with $AN$ as a free body, $NN'$ being = $dz$, $$\frac{pI}{c} = M \text{; whence } p' = \frac{Me}{I} \text{ . . . . . (3)}$$ $p$ and $p'$ are the respective normal stresses in the outer fibre in the transverse sections $N$ and $N'$ respectively. Now separate the block $MN'$, lying between these two consecutive sections, as a free body (in Fig. 249). And A diagram showing a block MN' separated into two parts, with arrows indicating the direction of shear force. Furthermore remove a portion of the top of the latter block, the portion lying above a plane parallel to the neutral surface, and consider this as a free body. This latter free body is shown in Fig. 250, with the system of forces representing the actions upon it of the portions taken away. The under surface, just laid bare, is a portion of a surface (parallel to the neutral surface) in which the above mentioned slipping, or shearing, tendency exists. The lower portion (of the block $MN'$) which is now removed exerted this A diagram showing a block MN' separated into two parts, with arrows indicating the direction of shear force. Fig. 249. Fig. 250. 286 MECHANICS OF ENGINEERING. rubbing, or sliding, force on the remainder along the under surface of the latter. Let the unknown intensity of this shearing force be $X$ (per unit of area); then the shearing force on this under surface is $-XY dx$, ($y$—see in figure, being the horizontal width of the beam at this distance $x'$ from the neutral axis of $N$) and makes its plane with the other forces of the system, which are the normal stresses between $\frac{F-c}{L-x'}$, and portions of $J$ and $F$, the respective total vertical shears. (The manner of distribution of $J$ over the vertical section is as yet unknown; see next article.) Putting $Z$ (horiz. compn.) = 0 in Fig. 350, we have $$\int_{x}^{x'} \frac{t}{e} p'dF - \int_{x'}^{\infty} \frac{p}{e} pdF - Xy' dx = 0$$ $$\therefore Xy' dx = \frac{p}{e} - \frac{p}{e} \int_{x'}^{\infty} dF$$ But from eqs. (1) and (2), $p' - p = (M - M') \frac{s}{T} dM,$ while from § 240 $dM = Jdx;$ $$\therefore Xy' dx = \frac{Jdx}{I} \int_{x'}^{\infty} dF; \therefore X = \frac{J}{I} \int_{x'}^{\infty} dF$$ (9) as the required intensity per unit of area of the shearing force in a surface parallel to the neutral surface and at a distance $x'$ from it. It is seen to depend on the "shear" $J$ and the moment of inertia $I$ of the whole vertical section; upon the horizontal thickness $y'$ of the beam at the surface in question; and upon the integral $\int_{x'}^{\infty} dF,$ which (from § 23) is the product of the area of that part of the vertical section extending from the surface in question to the outer fibers, by the distance of the centre of gravity of that part from the neutral surface. SHEAR IN FLEXURE. 287 It now follows, from § 209, that the intensity (per unit area) of the shear on an elementary area of the vertical cross section of a bent beam, and this intensity we may call $Z$, is equal to that $X$ just found, in the horizontal section which is at the same distance ($e$) from the neutral axis. 254. Mode of Distribution of $J$, the Total Shear, over the Vertical Cross Section.—The intensity of this shear, $Z$ (lbs. per sq. inch, for instance) has just been proved to be $$Z = X = \frac{J}{Ig} \int_{z}^{y'} zdF$$ (4) To illustrate this, required the value of $Z$ two inches above the neutral axis, in a cross section close to the abutment, in Ex. 5, § 252. Fig. 253 shows the position of the beam, it being as for the shaded portion lying above the locality in question, $y'$ 4 inches, and $\int_{z'}^{y'} zdF = (area)$ of shaded portion) × (distance of its centre of gravity from $X$,$A$) = (12.8 sq. in.) × (3.6 in.) = 46.08 cubic inches. The total shear $J$ is the abutment reaction = 600 lbs., while $I = \frac{1}{3}bh^2 = \frac{1}{3} \times 4 \times (10.4)^2 = 375$ biquad. inches. Both $J$ and $I$ refer to the whole section. $$Z = \frac{600 \times 46.08}{375 \times 4} = 18.42 \text{ lbs. per sq. in.},$$ quite insignificant. In the neighborhood of the neutral axis, where $z' = 0$, we have $y' = 4$ and $$\int_{z'}^{y'} zdF = \int_{0}^{4} zdF = 20.8 \times 2.6 - 54.8,$$ where $J$ and $I$ of course are the same as before. Hence for $z' = 0$ A diagram showing a beam with a shaded portion above a neutral axis. A diagram showing a cross-section of the beam with dimensions labeled. 288 MECHANICS OF ENGINEERING. $Z - Z_0 = 21.69$ lbs. per sq. in. At the outer fibre since $\int_{z'}^{z''} z'dF = 0$, $z'$ being $= e$, $Z$ is $= 0$ for a beam of any shape. For a solid rectangular section like the above, $Z$ and $z'$ bear the same relation to each other as the co-ordinates of the parabola in Fig. 253 (axis horizontal). Since in equation (4) the horizontal thickness, $y''$, from side to side of the section of the locality where $Z$ is desired, occurs in the denominator, and since $\int_{z'}^{z''} z'dF$ increases as $z'$ grows numerically smaller, the following may be stated, as to the distribution of $J$, the shear, in any vertical section, viz: The intensity (lbs. per sq. in.) of the shear is zero at the outer elements of the section, and for beams of ordinary shapes is greatest where the section crosses the neutral surface. For forms of cross section having thin webs its value may be so great as to require special investigation for safe design. Denoting by $Z_0$ the value of $Z$ at the neutral axis, (which $= X_0$ in the neutral surface where it crosses the vertical section in question) and putting the thickness of the substance of the beam = $b_0$ at the neutral axis, we have, $$Z_0 = X_0 \times \frac{J}{T_{b_0}} \times \left\{\begin{array}{l} {\text{area above neutral axis}}\\ {\text{(or below)}} \end{array}\right.\times \left\{\begin{array}{l} {\text{the dist. of its cent. grav.from that axis}} \end{array}\right.$$ (5) 255. Values of $Z_0$ for Special Forms of Cross Section.—From the last equation it is plain that for a prismatic beam the value of $Z_0$ is proportional to $J$, the total shear, and hence to the ordinate of the shear diagram for any particular ease of loading. The utility of such a diagram, as obtain- SHEAR IN FLEXURE. 289 ed in Figs. 234-237 inclusive, is therefore evident, for by locating the greatest shearing stress in the beam it enables us to provide proper relations between the loading and the form and material of the beam to secure safety against failure. The table in § 210 gives safe values which the maximum $Z_{e}$ in any case should not exceed. It is only in the case of beams with thin webs (see Figs. 238 and 240) however, that $Z_{e}$ is likely to need attention. For a Rectangle we have, Fig. 253, (see eq. 5, § 210) $$b_0 = b, I = \frac{1}{6}bh^3, \text{and } \int_{\text{sec}} zdF = \frac{1}{4}bh^2, \frac{1}{4}h = \frac{1}{4}bh^2$$ $$\therefore Z_e = X_e = \frac{1}{4} \cdot \frac{1}{4} \text{i.e., } = \frac{1}{4} (\text{total shear}) + (\text{whole area})$$ Hence the greatest intensity of shear in the cross-section is $\frac{1}{4}$ as great per unit of area as if the total shear were uniformly distributed over the section. Fig. 254 Fig. 255 Fig. 256 Fig. 257 For a Solid Circular section Fig. 254 $$Z_e = \frac{J}{Ih_0} \int_{\text{sec}} zdF = \frac{J}{\frac{1}{4}\pi r^3 \cdot 2r} \cdot \frac{4r}{2} = \frac{J}{\pi r^3}$$ (See § 26 Prob. 3). For a Hollow Circular section (concentric circles) Fig. 255, we have similarly, $$Z_e = \frac{J}{Ih_0} \int_{\text{sec}} zdF = \frac{J}{\frac{1}{4}\pi r^3 - \frac{1}{4}\pi r_0^3} \cdot \frac{4r - 4r_0}{2} = \frac{J}{\pi r^3 - \pi r_0^3}$$ 290 MECHANICS OF ENGINEERING. $$Z_{m} = \frac{J}{\pi (r_1^2 - r_2^2)} \left[ \frac{\pi r_1^3}{6} + \frac{\pi r_2^3}{6} \right]$$ Applying this formula to Example 2 § 253, we first have as the max. shear $J_m = J P = 1,735$ lbs., this being the abutment reaction, and hence (putting $x = (22 + 7)$) $$Z_m = \frac{4 \times 7 \times 1735(64 + 49.8)}{3 \times 39(256 - 150)(4 + 3.5)} = 294 \text{ lbs. per sq. in.}$$ which cast iron is abundantly able to withstand in shearing. For a Hollow Rectangular Beam, symmetrical about its neutral surface, Fig. 256 (box girders) $$Z_m = \frac{J (b_1 h_1^2 - b_2 h_2^2)}{\gamma m (b_1 h_1^2 - b_2 h_2^2) (b_1 - b_2)} = \frac{3}{2} \frac{J (h_1 h_1^2 - h_2 h_2^2)}{(b_1 h_1^2 - b_2 h_2^2) (b_1 - b_2)}$$ The same equation holds good for Fig. 257 (I-beam with square corners) but then $b_1$ denotes the sum of the widths of the hollow spaces. **256. Shearing Stress in the Web of an I-Beam.— It is usual to consider that, with I-beams (and box-beams) with the web vertical the shear $J$, in any vertical section, is borne exclusively by the web and is uniformly distributed over its section. That this is nearly true may be proved as follows, the flange area being comparatively large. Fig. 258. Let $F_s$ be the area of one flange, and $F_w$ that of the half web. Then since** A diagram showing a cross-section of an I-beam with a flange and a web. Fig. 256 $$I = \frac{1}{3} b_1 h_1^3 + 2 F_w \left( \frac{h_1}{3} \right)^3$$ SHEAR IN FLEXURE. 291 (the last term approximate, $\frac{1}{2}h_{0}$ being taken for the radius of gyration of $F_1$) while $$\int_{a}^{b}\frac{\mathrm{d}F}{F_1+\frac{F_2}{2}}=\frac{J}{h_0}, \quad \text{(the first term approx.) we have}$$ $$Z=\frac{J}{h_0}=\frac{J(\frac{1}{2}F_1+F_2)}{h_0}, \text{ which } =\frac{J}{h_0},$$ if we write $(3F_1+F_2)+(6F_1+2F_2)=\frac{5}{2}$. But $h_{0}$ is the area of the whole web, :. the shear per unit area at the neutral axis is nearly the same as if $J$ were uniformly distributed over the web. E. g., with $F_1=2$ sq. in., and $F_2=1$ sq. in. we obtain $Z=1.07(\frac{J}{h_0})$. Similarly, the shearing stress per unit area at $a$, the upper edge of the web, is also nearly equal to $J+h_{0}$ (see eq., 4 ($\S$54)) for then $$\left[\int_{-h_{0}}^{a}\frac{\mathrm{d}F}{F}\right]=F,\frac{1}{2}h_{0} \text{ nearly},$$ while $I$ remains as before. The shear per unit area, then, in an ordinary I-beam is obtained by dividing the total shear $J$ by the area of the web section. **Example**—It is required to determine the proper thickness to be given to the web of the 15-inch wrought-iron rolled beam of Example 3 of §252, the height of web being 13 inches, with a safe shearing stress as low as 4000 lbs. per sq. inch. N. J. Steel and Iron Co., (for bolts), the web being vertical. The greatest total shear, $A_{max}$ occurring at either support and being equal to half the load (see table §250) we have with $b_0=$ width of web, $$Z_{max}=\frac{J}{h_0}; \text{i.e. } 4000=\frac{13650}{b_0\times13}. \therefore b_0=0.26 \text{ inches}.$$ 292 MECHANICS OF ENGINEERING. (Units, inch and pound). The 15-inch light beam of the N. J. Co. has a web $\frac{3}{4}$ inch thick, so as to provide for a shear double the value of that in the foregoing example. In the middle of the span $Z_0 = 0$, since $J = 0$. 257. Designing of Riveting for Built Beams.—The latter are generally of the I-beam and box forms, made by riveting together a number of continuous shapes, most of the material being thrown into the flange members. E.g., in fig. 259, an I-beam is formed by riveting together, in the manner shown in the figure, a "vertical stem plate" or web, four "angle-irons," and two "flange-plates," each of Figure 259 Figure 260 these seven pieces being being continuous through the whole length of the beam. Fig 260 shows a box-girder. If the riveting is well done, the combination forms a single rigid beam whose safe load for a given span may be found by foregoing rules; in computing the moment of inertia, however, the portion of cross section cut out by the rivet holes must not be included. (This will be illustrated in a subsequent paragraph.) The safe load having been computed from a consideration of normal stresses only, and the web being made thick enough to take up the max. total shear, it is evident that no further design is required, for through whose agency the web and flanges are caused to act together as a single continuous rigid mass. It will be on the side of safety to consider that at a given SHEAR IN FLEXURE. 293 locality in the beam the shear carried by the rivets connecting the angles and flanges, per unit of length of beam, is the same as that carried by those connecting the angles and the web (" vertical stem-plate"). The amount of this shear may be computed from the fact that it is equal to that occurring in the surface (parallel to the neutral surface) in which the angle joint is located, i.e., in the web and flanges of continuous substance, as in a solid I-beam. But this shear must be of the same amount per horizontal unit of length as it is per vertical linear unit in the web itself, where it joins the flange ; (for from $254 Z = X$.) But the shear in the vertical section of the web, being uniformly distributed, is the same per vertical linear unit at the junction with the flange as at any other part of the web section ($\S 256$), and the whole shear on the vertical section of web = $J$, the " total shear" of that section of the beam. Hence we may state the following : The riveting connecting the angles with the flanges, (or the web with the angles) in any locality of a built beam, must safely sustain a shear equal to $J$ on a horizontal length equal to the height of web. The strength of the riveting may be limited by the resistance of the rivet to being sheared (and this brings into account its cross-section) upon the crushing resistance of the side of the rivet hole in the plate (and this involves both the diameter of the rivet and the thickness of the metal in the web, flange, or angle). In its practice the N. J. Steel and Iron Co. allows 7500 lbs. per sq. inch shearing stress in the rivet (wrought iron), and 12500 lbs. per sq. inch compressive resistance in the side of the rivet-hole, the axial plane section of the hole being the area of reference. In $\S 259$ the rivets connecting the web with the angles are in double shear, which should be taken into account in considering their shearing strength, which is then double : those connecting the angles and flange plates are in A diagram showing a beam with angles and flanges connected by rivets. 294 MECHANICS OF ENGINEERING. single shear. In fig. 260 (box-beam) where the beam is built of two webs, four angles, and two flange plates, all the rivets are in single shear. If the web plate is very high or thick with its thickness, vertical stiffeners in the form of T iron may need to be riveted upon them laterally (see § 314). **Example**—A built I-beam of wrought iron (see fig. 259) is to support a uniformly distributed load of 40 tons, its extremities resting on supports 20 feet apart, and the height and thickness of web being 20 ins. and $\frac{1}{2}$ in. respectively. How shall the rivets, which are $\frac{1}{4}$ in. in diameter, be spaced between the web and the angles which are also $\frac{1}{2}$ in. in thickness? Referring to fig. 259 we find that $f$ = 0.008 sq. inches at the top and bottom, diminishing regularly to zero at the middle, where no riveting will therefore be required. (Units inch and pound). Near a support the riveting must sustain for each inch of length of beam a shearing force of $(f \times$ height of web) = 40000 + 20 ins. = 2000 lbs. Each rivet, having a sectional area of $\frac{1}{4} \pi (\frac{1}{4})^2 = 0.00$ sq. inches, can bear a shearing of 0.60 X 7500 = 4500 lbs. in single shear, and $\frac{1}{4}$ of 9000 lbs. in double shear, which is the present case. But the safe compressive resistance of the side of the riveted plate in either the web or angle is only $\frac{1}{4}$ in. X $\frac{1}{4}$ in. X 12500 = 5470 lbs., and thus determines the spacing of the rivets as follows: 2000 lbs. + 5470 gives 0.36 as the number of rivets per inch of length of beam, i.e., they must be $1 \div 0.36 = 2.7$ inches apart, centre to centre, near the supports; 5.4 inches apart at $\frac{1}{4}$ the span from a support; none at all in the middle. However, "the rivets should not be spaced closer than $2\frac{1}{4}$ times their diameter, nor farther than 16 times the thickness of the plate they connect," is the rule of the N. J. Co. As for the rivets connecting the angles and flange plates, being in two rows and opposite (in pairs) the safe shearing force is A diagram showing a box-beam structure with various components labeled. FLEXURE, BUILT BEAM. 295 ing resistance of a pair (each in single shear) is 9,000 lbs., while the safe compressive resistance of the sides of the two rivet holes in the angle iron (the flange plate being much thicker) is = 10,940 lbs. Hence the former figure (9,000) divided into 2,000 lbs., gives 0.25 as the number of pairs of rivets per inch of length of the beam; i.e., the rivets in one row should be spaced 4.5 inches apart, centre to centre, and the spacing between rows must be inversely in inverse ratio to the distance from the middle of spans, (bearing in mind the practical limitation just given). If the load is concentrated in the middle of the span, instead of uniformly distributed, $J$ is constant along each half-span, (see fig. 234) and the rivet spacing must accordingly be made the same at all localities of the beam. SPECIAL PROBLEMS IN FLEXURE. 258. Designing Cross Sections of Built Beams.—The last paragraph dealt with the riveting of the various plates; we now consider the design of the plates themselves. Take for instance a built I-beam, fig. 261 ; one vertical stem- A diagram showing a cross-section of a built I-beam. The top part shows a vertical stem with a horizontal plate on top. Below this, there are two horizontal plates connected by vertical bars. The bottom part shows another vertical stem with a horizontal plate on top. Fig. 261. 261 296 MECHANICS OF ENGINEERING. plate, four angle irons, (each of sectional area = A, remaining after the holes are punched, with a gravity axis parallel to, and at a distance = a from its base), and two flange plates of width = b, and thickness = t. Let the whole depth of girder = h, and the diameter of a rivet hole = d. To see readily that the forces are conservative forces induced in this section by $M_c$ in-line ($M_c$ being the greatest moment in the beam which is prismatic) we have from § 239, $$M_c = \frac{R^2T}{h} \tag{1}$$ For wrought iron = 12,000 lbs. per sq. inch, $c$ is $\frac{3}{4}$ $h$ while $I$, the moment of inertia of the compound section, is obtained as follows, taking into account the fact that the rivet holes cut out part of the material. In dealing with the sections of the angles and flanges, we consider them concentrated at their centres of gravity (an approximation, of course,) and treat their moments of inertia about $N$ as single terms in the series $\int dF^2$ (see § 85). The subtractive moments of inertia for the rivet holes in the web are similarly expressed ; let $t_b =$ thickness of web. $$\therefore I_0 \text{ for web } = \frac{1}{4}b_h(h-2t)-2bt^2(\frac{3}{4}-a)^2$$ $$\therefore I_4 \text{ for four angles } = 4A(\frac{3}{4}-a)^2$$ $$\therefore I_6 \text{ for two flanges } = 2(2t-F)(\frac{3}{4})^2$$ the sum of which makes the $I_6$ of the girder. Eq. (1) may now be written $$\frac{M_cA}{2E} = I_6 \tag{2}$$ which is available for computing any one unknown quantity. The quantities concerned in $I_4$ are so numerous and they are combined in so complex a manner that in any numerical example it is best to adjust the dimensions of the section to each other by successive assumptions and FLEXURE BUILT BEAM. 297 trials. The size of rivets need not vary much in different cases, nor the thickness of the web-plate, which as used by the N. J. Co. is "rarely less than $\frac{3}{4}$ or more than $\frac{5}{8}$ inch thick." The same Co. recommends the use of a single size of angle iron, viz., $5 \times 3 \times \frac{3}{4}$", for built girders of heights ranging from 12 to 36 inches, and also $\frac{3}{4}$ in. rivets, and gives tables computed from eq. (2) for the proportionate strength of each portion of the compound section. EXAMPLE.--(Units, inch and pound). A built I-beam with end supports, of span = 20 ft. = 240 inches, is to support a uniformly distributed load of 36 tons = 72,000 lbs. If $\frac{3}{4}$ inch rivets are used, angle irons $3 \times 3 \times \frac{3}{4}$", vertical web $\frac{3}{4}$" in thickness, and plates 1 inch thick for flanges, how wide ($b$ ?) must these flange-plates be ? taking $h = 22$ inches = total height of girder. Solution.--From the table in § 250 we find that the max. $M$ for this case is $\frac{3}{4} W$, where $W =$ the total distributed load (including the weight of the girder), and $l =$ span. Hence the left hand member of eq. (2) reduces to $$\frac{P T}{16} \cdot \frac{A}{W} = \frac{72000 \times 240 \times 22}{16 \times 12000} = 1980$$ That is, the total moment of inertia of the section must be $= 1980$ biquad. inches, of which the web and angles supply a known amount, since $a = \frac{3}{4} l$, $t = \frac{3}{4} h$, $c' = 1$ inch, but $b' - c' = h - t = 25$ inches are known, while the remainder must be furnished by the flanges, thus determining their width $b$, the unknown quantity. The effective area, $A$, of an angle iron is found thus: The full sectional area for the size given, $- 3 \times \frac{3}{4} + 2 \frac{3}{4} \times \frac{3}{4} = 2.75$ sq. inches, from which deducting for two rivet holes we have $$A = 2.75 - 2 \times \frac{3}{4} \times \frac{3}{4} = 2.0$$ sq. in. The value $a = 0.90$ is found by cutting out the shape 298 MECHANICS OF ENGINEERING. of two angles from sheet iron, thus : and balancing it on a knife edge. (The gaps left by the rivet holes may be ignored, without great error, in finding $a$). Hence, substituting we have $$I_1 \text{ for web} = \frac{1}{3} \cdot \frac{1}{2} \times 30^2 - 2 \times \frac{1}{2} \cdot \frac{1}{2} \cdot (8 \cdot 5)^2 = 282.3$$ $$I_2 \text{ for four angles} = 4 \times 2 \times (31.0)^2 = 662.5$$ $$I_3 \text{ for two flanges} = 2(b-\frac{1}{2}) \times 1 \times (10)^2 = 220.4(b-1.5)$$ :. $1980 = 282.3 + 662.5 + (b-1.5)290.4$ whence $b = 4.6 + 1.5 = 6.1$ inches the required total width of each of the 1 in. flange plates. This might be increased to 6.5 in so as to equal the united width of the two angles and web. The rivet spacing can now be designed by § 257, and the assumed thickness of web, $\frac{1}{4}$ in., tested for the max. total shear by § 256. The latter test results as follows: The max. shear $J_a$ occurs near either support and $= \frac{1}{4}W - 36,000\text{ lbs}.$. , calling $b'$ the least allowable thickness of web in order to keep the shearing stress as low as 4,000 lbs. per sq. inch, $$b'_{a} \times 30^2 \times 4000 - 36000 : b'_{a} = 0.45''$$ showing that the assumed width of $\frac{1}{4}$ in. is safe. This girder will need vertical stiffeners near the ends, as explained subsequently, and is understood to be supported laterally. Built beams of double web, or box-form, (see Fig. 390) do not need this lateral support. **259. Set of Moving Leads.—When a locomotive passes over a number of parallel prismatic girders, each one of which experiences certain detached pressures corresponding to the different wheels, by selecting any definite position of the wheels on the span, we may easily compute the reactions of the supports, then form the shear diagram, and finally as in § 243 obtain the max. moment, $M_m$, and the A diagram showing two angles connected by a rivet hole. Fig. MII. FLEXURE MOVING LOADS. 299 max. shear $J_m$ for this particular position of the wheels. But the values of $M_n$ and $J_n$ for some other position may be greater than those just found. We therefore inquire which will be the greatest moment among the infinite number of ($M_n$)'s (one for each possible position of the wheels on the span). It is evident from Fig. 238 from the nature of the moment diagram, that when the pressures or loads are detached, the $M_n$ for any position of the loads, which occur under one wheel, must occur under one of the loads (i.e., under a wheel). We begin . by asking: What is the position of the set of moving loads when the moment under a given wheel is greater than will occur under that wheel in any other position? For example, in Fig. 269, in what position of Diagram showing a set of wheels with moving loads. Fig. 269. loads $P_1$, $P_2$, etc. on the span will the moment $M_n$, i.e., under $P_1$ be a maximum as compared with its value under $P_1$ in any other position on the span. Let $R$ be the resultant of the loads which are now on the span, its variable distance from $O$ to $A$ being $\bar{x}$, and let $P_1 = P$, while $a$, $b$, etc., are the fixed distances between the loads (wheels). For any values of $\bar{x}$, as the loading moves through the range of motion within which no wheel of the set under consideration goes off the span, and no new wheel comes on it, we have $\bar{x} = \frac{z}{l}$ $E$, and the moment under $P_1$ $$-M_n = R_1(\bar{x} - a) - P_1 - P(b + c)$$ i.e. $$M_n = \frac{P}{l}(\bar{x} - b + c) - P_1 - P(b + c)$$ (1) 300 MECHANICS OF ENGINEERING. In (1) we have $M$ as a function of $\bar{x}$, all the other quantities in the right hand member remaining constant as the loading moves; $\bar{x}$ may vary from $x-a+d$ to $x-l-(c+b-d')$. For a max. $M$, we put $dM/d\bar{x}=0$, i. e. $$\frac{R}{l}(-2\bar{x}+a)=0 \therefore \bar{x} \text{(for Max } M) = \frac{a}{2} + \frac{c}{2}$$ (For this, or any other value of $\bar{x}$, $d^2M/d\bar{x}^2$ is negative, hence a maximum is indicated). For a max. $M$, then, $R$ must be as far ($\frac{a}{2}$) on one side of the middle of the span as $P_1$ is on the other; i.e., as the loading moves, the moment under a given wheel becomes a max. when that wheel and the centre of gravity of all the loads (then on the span) are equidistant from the middle of the span. In this way in any particular case we may find the respective max. moments occurring under each of the wheels during a passage, and the greatest of these is the $M_0$ to be used in the equation $M = R'f + c$ for safe loading.* As to the shear $J$, for a given position of the wheels this will be the greatest at one or the other support, and equals the reaction at that support. When the load moves toward either support the shear at that end of the beam evidently increases so long as no wheel rolls completely over and beyond it. To find J max., then, dealing with each successive turn, we take those positions at the support where the loading is successively so placed that consecutive wheels, in turn, are on the point of rolling off the girder at that end; the greatest of these is the max. shear, $J_0$. As the max. moment is apt to come under the heaviest load it may not be necessary to deal with more than one or two wheels in finding $M_0$. **Example:** Given the following wheel pressures, $$A < 8 \ldots B > 5 \ldots C < 4 \ldots D > 3$$ 4 tons 5 tons 6 tons 7 tons on one rail which is continuous over a girder of 20 ft. span, under a locomotive. *Since this may be regarded as a case of "sudden application" of a load, it is customary to make $R'$ much smaller than for a dead load from one-third to one-half smaller. FLEXURE MOVING LOADS. 301 1. Required the position of the resultant of $A$, $B$, and $C$; 2. " " $A$, $B$, $C$, and $D$; 3. " " " " $B$, $C$, and $D$. 4. In what position of the wheels on the span will the moment under $B$ be a max.? Ditto for wheel $C$? Required the value of these moments and which is $M_{m}$? 5. Required the value of $J_{m}$ (max. shear), its location and the position of loads. Results.—(1) 7.8° to right of $A$. (2) 10° to right of $A$. (3) 4.4° to right of $B$. (4) Max. $M_{m} = 1,273,000$ inch lbs. with all the wheels on; Max. $M_{m} = 1,440,000$ inch lbs. with wheels $B$, $C$, and $D$ on. (5) $J_{m} = 13.6$ tuns at right support with wheel on, but the support is not at right. 280. Special Eccentric Load. In the following special cases of prismatic beam, peculiar in the distribution of the loads, or mode of support, or both, the main objects sought are the values of the max. moment $M_{m}$ for use in the equation $$M_m = \frac{P_l}{R^2}$$ (see §239); and of the max. shear $J_{m}$, from which to design the web riveting in the case of an I or box-girder. The modes of support will be such that the reactions are independent of the form and material of the beam (the weight of beam being neglected). As before, the beam may be slight, and the forces are all perpendicular to the beam. The present problem is that in fig. 285, the beam being prismatic, supported at the ends, with a single eccentric load, $P$. We shall first disregard the weight of the beam itself. Let the span = $l_1 + l_2$. First considering the whole beam free we have the reactions $R_1 = P_l + l$ and $R_2 = P_l + l$ Making a section at m and having Om free, x being < $l_1$, z (vert compn.) = 0 gives 302 MECHANICS OF ENGINEERING. $R_t - J = 0$, i.e., $J = R_t$; while from $I (mom.)_{=0}$ we have $$\frac{pL}{e} \cdot R_p = 0 \cdot M = R_p \cdot \frac{P_h}{l}$$ These values of $J$ and $M$ hold good between $0$ and $C$, $\omega$ being constant, while $M$ is proportional to $x$. Hence for $OC$ the shear diagram is a rectangle and the moment diagram a straight line. The max. $M$ for $OC$ is for $x = l_0$ and is $P_l l_0 + l$. This is the max. $M$ for the beam, since between $C$ and $B$, $M$ is proportional to the distance of the section from $B$. :. $M_m = \frac{P_l l_0}{t}$ and $\frac{R_f}{e} = \frac{P_l l_0}{t}$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) is the equation for safe loading. $J = R_t$, in any section along $CB$, and is opposite in sign to what it is on $OC$; i.e., practically, if a dove-tail joint existed anywhere on $OC$ the portion of the beam on the right of such section would slide downward relatively to the left, and thus cause failure at $CB$. Evidently the max. shear $J_m = R_t$ or $R_s$ as $l_t$ or $l_r$ is the greater segment. It is also evident that for a given span and given beam the safe load $P_r$, as computed from eq. (1) above, becomes very large as its point of application approaches a support; this would naturally be expected but not without limit, as the shear for sections between the load and the support is equal to the reaction at the near support and may therefore be considered as due to the weight of the web or the spacing of the rivets, if any, is considered. Secondly, considering the weight of the beam, or any uniformly distributed loading, weighing w lbs. per unit of length of beam, in addition to P, Fig. 264, we have the reactions $$R_t = \frac{P_h}{t} + \frac{W}{2}; \text{and } R_s = \frac{P_h}{t} + \frac{W}{2}$$ Let $l_t > l_r$; then for a portion $OM$ of length $z < l_0$, moments about m give FLEXURE. SPECIAL PROBLEMS. 303 $$\frac{P}{I} - R_x + wx \cdot \frac{x}{2} = 0$$ i.e., on $OC$, $M = R_x x - \frac{1}{2} wx^2$ Evidently for $x = 0$ (i.e. at $O$) $M = 0$, while for $x = l_1$ (i.e. at $C$) we have, putting $w = W + I$, $$M_c = R_{c1} - \frac{1}{2} w l_1^2 = \frac{P l_1}{l} + \frac{W l_1}{2} - \frac{W P}{2}$$ It remains to be seen whether a value of $M$ may not exist in some section between $O$ and $C$, i.e., for a value of $x < l_1$ in eq. (2)), still greater than $M_c$. Since (2) gives $M$ as a continuous function of $x$ between $O$ and $C$, we put $\frac{dM}{dx} dz = 0$, and obtain, substituting the value of the constants $R_x$ and $w$, $$R_x - wx = 0 \therefore z_c$$ for $M > M_c$ $$\frac{P l_1}{l} + \frac{W l_1}{2} - \frac{W P}{2}$$ (4) This must be for $M_{max}$, since $\frac{dM}{dx} dz$ is negative when this value of $x$ is substituted. If the particular value of $x$ given by (4) is $z_c$, the corresponding value of $M$ will be less than that from (2) will occur on $OC$ and will be greater than $M_c$. (Diagrams II. in fig. 264 show this case); but if $x > z_c$, we are not concerned with the corresponding value of $M$, and the greatest $M$ on $OC$ would be $M_c$. For the short portion $BC$, which has moment and shear diagrams of its own continuous with those for $OC$, it may easily be shown that $M_c$ is the greatest moment of any section. Hence the $M$ Diagram showing a flexure problem with sections O, C, and B. Fig. 264 304 MECHANICS OF ENGINEERING. max., or $M_{\infty}$, of the whole beam is either $M_c$ or $M_a$, according as $x_c$ is > or < $l_c$. This latter criterion may be expressed thus, (with $L = \frac{1}{2} l$ denoted by $l_a$ the distance of $T$ from the middle of the span): From (eq. 4) $\left[\frac{P_l}{W} + \frac{1}{2} l\right] > \left[\frac{P}{W}\right]$ is equivalent to $\left[\frac{P}{W} < \frac{l_a}{2}\right]$ and since from (4) and (2) $$M_c = \left[\frac{P_l}{T + \frac{1}{2} W}\right]\left[\frac{P_l}{T + \frac{1}{2} W}\right] - \frac{1}{2} W\frac{P_l}{T + \frac{1}{2} W}$$ (5) The equation for safe loading is $$\frac{R_f}{e} = M_c, \text{ when } P_W > \frac{l_1}{l_a}$$ and $$\frac{R_f}{e} = M_a, \text{ when } P_W < \frac{l_1}{l_a}$$ See eqs. (3) and (5) If either $P$, $W$, $l_a$ or $l$, is the unknown quantity sought, the criterion of (6) cannot be applied, and we use both equations in (6) and then discriminate between the two results. The greatest shear is $J_m = B_1$, in Fig. 264, where $t$ is > $l_a$. 261. Two Equal Terminal Loads. Two Symmetrical Supports Fig. 265. (Same case as in Fig. 231, p. 388). Neglect weight of beam. The reaction at each support being $-P$, (from symmetry), we have for a free body $Om$ with $x < l_c$ $$Px - P(x-l_1) = Mz = Pz$$ (1) while where $x > l_c$ and $< l_1 + l_a$ $$Px - P(x-l_1) = Mz = Pl$$ (2) That is, see (1), $M$ varies directly with $x$ between $O$ and $C$, while between $C$ and $D$ it is constant. Hence for safe loading $$\frac{R_f}{e} = M_c \text{ i.e. } \frac{R_f}{e} = Pl$$ (3) A diagram showing two symmetrical supports with equal terminal loads. FLEXURE. SPECIAL PROBLEMS. 305 Diagram showing a beam with flanges and web. Fig. 265. The construction of the moment diagram is evident from equations (1) and (2). As for $J$, the shear, the same free bodies give, from $\Sigma$ (vert. forces) = 0. On $O C \cdot J = P \quad \ldots \quad (4)$ On $C D \cdot J = P - P = zero \quad (5)$ (4) and (5) might also be obtained from (1) and (2) by writing $J = d M + dx$, but the former method is to be preferred in most cases, since the latter requires $M$ to be expressed as a function of $x$ while the former is applicable for examining separate sections without making use of a variable. If the beam has a flange, the fact that $J$ is zero anywhere on $CD$ would indicate that we may dispense with a web along $CD$ to unite the two flanges; but the lower flange being in compression and forming a "long column" would tend to buckle out of a straight line if not stayed by a web connection with the other, or some equivalent bracing. 282. Uniform Load Over Part of the Span. Two End Supports Fig. 266. Let the load $W$, extending from one support over a portion $c$ of the span, (on the left, say,) so that $W = wc$, $w$ being the load per unit of length. Neglect weight of beam. For a free body $Om$ of any length $x < OD$ (i.e. < c), 2 moments = 0 gives $$\frac{P}{e} + \frac{wx^2}{2} - R_x = 0 : M - R_x x = \frac{wx^2}{2} \quad \ldots \quad (1)$$ which holds good for any section on $OB$. As for sections on $BC$ it is more simple to deal with the free body $m'c$, of length $x' < CB$ from which we have $M - R_x x' = \ldots \quad (2)$ 306 MECHANICS OF ENGINEERING. which shows the moment curve for $BC$ to be a straight line $DC$, tangent at $D$ to the parabola $O'D$ representing eq. (1). (If there were a concentrated load at $B$, $CD$ would meet the tangent at $D$ at an angle instead of coinciding with it; let the student show why, from the diagram.) The shear for any value of $x$ on $OB$ is: On $OB \dots J = R_x - wx \dots \dots (3)$ while on $BC \dots J = R_p \text{ constant} \dots (4)$ The shear diagram is constructed accordingly. To find the position of the max. ordinate of the parabola, (and this from previous statements concerning the tangent at the point $D$ must occur on $OB$, as will be seen and will . be the $M_o$ for the whole beam) we put $J = 0$ in eq (3) whence $x (\text{for } M_o) = \frac{R_1}{w} - \frac{W(l-1)}{l} = c - \frac{e^2}{2l} \dots (5)$ and is less than $c$, as expected. [The value of $R_1 = W(l-1)$, $\frac{e^2}{2l}(wx + l)(l-1), (\text{the whole beam free})$ has been substituted.] This value of $x$ substituted in eq. (1) gives $M_o = (1-\frac{e^2}{4l})\frac{e^2}{4l}, \frac{e^2}{4l}. Wc, \dots \frac{RI}{e} = \frac{e^2}{4l}(1-\frac{e^2}{4l})\frac{e^2}{4l}J^1Wc \dots (6)$ is the equation for safe loading. The max. shear $J_o$ is found at $O$ and is = $B_o$ which is evidently > $B_o$, at $C$. FLEXURE. SPECIAL PROBLEMS. 307 **263. Uniform Load Over Whole Length With Two Symmetric Supports.** Fig. 387.—With the notation expressed in the figure, the following results may be obtained, after having divided the length of the beam into three parts for separate treatment, as indicated by the external forces, which are the distributed load $W$, the two reactions each $\frac{1}{2} W$. The moment curve is made up of parts of three distinct parabolas, each with its axis vertical. The central parabola may sink below the horizontal axis of reference if the supports are too far apart, in which case (see Fig.) the vertical curve of the beam itself becomes concave upward between the points $E$ and $F$ of "contrary flexure". At each of these points the moment must be zero, since the radius of curvature is $\infty$ and $M = EI + \rho$ (see § 231) at any section; that is, at these points the moment curve crosses its horizontal axis. As to the location and amount of the max. moment $M_a$, inspecting the diagram we see that it will be either at $H_1$ the middle, or at both of the supports $B$ and $C$ (which from symmetry have equal moments), i.e., with $l$ total length), $$M_a \left[ \text{and}: \frac{R_f l}{e} = \begin{cases} \text{either } \frac{WF}{2l}(L^2 - L_1^2) & \text{at H} \\ \frac{WF}{2l} & \text{at B and C} \end{cases} \right.$$ according to which is the greater at any given case; i.e., according as $l_1$ is > or < $\sqrt{\frac{8}{\rho}}$. The shear close on the left of $B = wL_1$, while close to the right of $B$ it is $\frac{1}{2} W L_1$. (It will be noticed that in this case since the load only goes beyond the support, the shear near the support is not equal to the reaction there, as it was in some preceding cases.) A diagram showing a beam with two symmetric supports and a uniform load over its entire length. 308 MECHANICS OF ENGINEERING. Hence $J_m = \left\{ \begin{array}{l} w_0 \\ w_0 - w_{\text{surf}} \end{array} \right.$ according as $z \geq z_0$ 264. Hydrostatic Pressure Against a Vertical Plank.—From elementary hydrostatics we know that the pressure, per unit area, of incipient water against the vertical side of a tank, varies directly with the depth, $x$, below the surface, and equals the weight of a prism of water whose altitude $= x$, and whose sectional area is unity. See Fig. 268. A diagram showing a vertical plank with a triangular section. The top of the plank is labeled "O", the bottom is labeled "B". The sides of the triangle are labeled "a", "b", and "c". The base of the triangle is labeled "m". The height of the triangle is labeled "h". The width of the plank at the top is labeled "w_0". The width of the plank at the bottom is labeled "w_0 - w_{\text{surf}}". The depth of the water below the surface is labeled "x". The weight of the water is labeled "w_0". The weight of the water at the surface is labeled "w_0 - w_{\text{surf}}". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at the top is labeled "w_0". The weight of the water at the bottom is labeled "w_0 - w_{\text{surf}}". The weight of the water at FLEXURE. SPECIAL PROBLEMS. 309 (From (1) when $x = l$, we have for the total water pressure on the beam $W_x = \frac{y}{g} \cdot \frac{P}{l}$ and since one-third of this will be borne at $O$ we have $R_x = \frac{y}{g} \cdot \frac{P}{l}$.) Now putting $\zeta$ moma. about the neutral axis of $m = 0$, for, $om$ free, we have $$R_{px} - W_x \cdot \frac{P}{l} = 0 \cdot M = \frac{y}{g} \cdot P(x-x) \dots \dots \quad (2)$$ (which holds good from $x = 0$ to $x = l$). From $\Sigma$ (horiz. forces) = 0 we have also the shear $$J = R_x - W_x = \frac{y}{g} \cdot P(l-3x) \dots \dots \quad (3)$$ as might also have been obtained by differentiating (2), since $J = dM/dx$. By putting $J = 0$ ($\S$ 240, corollary) we have for a max. $M_x = t + \sqrt{3}$, which is less than $l$ and hence is applicable to the problem. Substitute this in eq. 2, and reduce, and we have $$\frac{P}{e} = H_{xx}, i.e. \frac{P}{e} = g \cdot \sqrt{\frac{l}{3}} \cdot r e \dots \dots \quad (4)$$ as the equation for safe loading. 265. Example.--If the thickness of the plank is $h$, required is $t$, if $R'$ is taken = 1000 lbs. per sq. in. for timber ($\S$ 251), and $l = 6$ feet. For the inch-pound-second system of units, we must substitute $R' = 1000$ lbs., $72$ inches; $r = 0.008$ lbs. per cubic inch; the heaviness of water in this case is $\frac{1}{2}$ lb. per cubic inch while $I = h^2t + 12$, ($\S$ 247), and $e = \frac{1}{4}h$. Hence from (4) we have $$1000 h^2t = 0.0365 \times 72^2, \therefore h = 5.16, . . . h = 2.27\text{ in}.$$ It will be noticed that since $x$ for $M_x = t + \sqrt{3}$, and not $\frac{1}{4}l$, $M_x$ does not occur in the section opposite the resultant of the water pressure; see Fig. 268. The shear curve is a parabola here; eq. (3). A diagram showing a beam with a load applied at various points along its length. 310 MECHANICS OF ENGINEERING. 268. The Four x-Derivatives of the Ordinate of the Elastic Curve —if $y = f_{\text{func}}(x)$ is the equation of the elastic curve for any portion of a loaded beam, on which portion the load per unit length of the beam is $w$—either zero, (Fig. 254) or constant, (Fig. 255)—we can prove, as follows, that $w$ is the x-derivative of the shear. Fig. 269. Let $N$ and $N'$ be two consecutive cross-sections of a loaded beam, and let the block between them, bearing its portion, $wdx$, of a distributed load, be considered free. The elastic forces consist of the two stress-couples (tensions and compressions) and the two shears, $J$ and $\frac{dJ}{dx}$ being the shear-increment consequent upon receiving its increment $dz$. By putting $z$ (vert. component) = 0 we have: $$J + dJ - w dx - J = 0 \therefore w = \frac{dJ}{dx}$$ Q. E. D. But $J$ itself = $dM + dx$, (§ 240) and $M = (dfy + dx^2)EL$. By substitution, then, we have the following relations: $$y = f_{\text{func}}(x) = \text{ordinate at any point of the elastic curve} \quad (1)$$ $$\frac{dy}{dx} = a = \text{slope at any point of the elastic curve} \quad . . . \quad (2)$$ $$EI \frac{d^2y}{dx^2} = M = \text{ordinate (to scale) of the moment curve} \quad (3)$$ $$EI \frac{d^2y}{dx^2} = \text{the shear}, J = \left\{\begin{array}{l}\text{the ordinate (to scale)} \\ \text{of the shear diagram}\end{array}\right\} \quad . . . \quad (4)$$ $$EI \frac{d^2y}{dx^2} = w = \left\{\begin{array}{l}\text{the load per unit of length} \\ \text{of beam at ordinate (to scale)} \\ \text{of curve of loading}\end{array}\right\} \quad . . . \quad (5)$$ It, then, the equation of the elastic curve (the neutral line of the beam itself); a reality, and not artificial like the **FLEXURE SPECIAL PROBLEMS.** 311 other curves spoken of) is given; we may by successive differentiation, for a prismatic and homogeneous beam so that both $E$ and $I$ are constant, find the other four quantities mentioned. As to the converse process, i.e. having given $w$ as a function of $x$, to find expressions for $J$, $M$ and $y$ as functions of $x$ this is more difficult, since in taking the $x$-anti-derivative, an unknown constant must be added and determined. The problem just treated in § 264, however, offers an exception to this case since $w$ is the same function of $x$, along the whole beam, so there is therefore but one elastic curve to be determined. We . . begin, numbering backward, with $$EI \frac{dy}{dx} = -r \frac{dw}{dx} \quad (5a)$$ [Note: This derivative ($dJ + dz$) is negative since $dJ$ and $dz$ have contrary signs.] But writing out this equation for $z=0$, i.e. for the point $O$, where the shear $R_0$ we have $R_0 = 0 + Const.$ . . . Const. = $R_0$ and hence write $$EI \frac{dy}{dx} = -r \frac{dw}{dx} + R_0 \quad (Shear) \quad (4a)$$ Again taking the $x$-anti-derivative of both sides $$\text{(Moment)} = EI \frac{dy}{dx} = -r \frac{dw}{dx} + R_0 + Const. = 0 \quad (3a)$$ [At $O$, $x=0$ also $M_0 = Const. = 0$.] Again, $$EI \frac{dy}{dx} = -r \frac{dw}{dx} + C'$$ At $O$, where $x=0 dy+dz=a_0=$ the unknown slope of the elastic line at $O$, and hence $C'=EI a_0$ $$EI \frac{dy}{dx} = -r \frac{dw}{dx} + a_0 + EI a_0 \quad (2a)$$ 312 MECHANICS OF ENGINEERING. Passing now to $y$ itself, and remembering that at $O$, both $y$ and $x$ are zero, so that the constant, if added, would—zero, we obtain (inserting the value of $R_0$ from last §) $$Ely = -\gamma b \frac{x^2}{120} + yb \frac{dx^2}{36} + Efa_x \quad . \quad (1a)$$ the equation of the elastic curve. This, however, contains the unknown constant $\alpha$,= the slope at $O$. To determine $\alpha$, write out eq. (1a) for the point B, Fig. 268, where $x$ is known to be equal to $l$, and $y$ to be = zero, solve for $\alpha$, and insert its value both in (1a) and (2a). To find the point of max. $y$ (i.e., of greatest deflection) in the elastic curve, write the equation, i.e., $dy/dx = 0$ (see § 5), and solve for four values which will be obtained by which the one lying between 0 and l is obviously the one to be taken. This value of $x$ substituted in (1a) will give the maximum deflection. The location of this maximum deflection is neither at the centre of action of the load ($x = \frac{2}{3} l$) nor at the section of max. moment ($x = l - \sqrt{\frac{3}{3}}$). The qualities of the left hand members of equations (1) to (5) should be carefully noted. E.g., in the inch-pound-second system of units we should have: 1. $y$/a linear quantity = (so many) inches. 2. $dy/dx$ (an abstract number) = (so many) abstract units. 3. $M$/a moment = (so many) inch-pounds. 4. $J$/a shear, i.e., force = (so many) pounds. 5. $w$/force per linear unit = (so many) pounds per running inch of beam's length. As to the quantities $E$, and $I$, individually, $E$ is pounds per sq. in., and $I$ has four linear dimensions, i.e. (so many) bi-quadratic inches. PLEXURE SPECIAL PROBLEMS. 313 287. Resilience of Beam with End Supports.—Fig. 270. If a mass whose weight is $G$ (a large compared with that of beam) be allowed to fall freely through a height $h$ upon the centre of a beam supported at its extremities, the pressure $P$ felt by the beam increases from zero at the first instant of contact to a maximum $F_{m}$ as already stated in §236, in which the equation was derived, $d_p$ being small compared with $h$, $$Gh = \frac{1}{96} \cdot \frac{P_{m} P}{E I} \cdot \ldots \cdot (a)$$ The elastic limit is supposed not passed. In order that the maximum normal stress in any outer fibre shall at most be $R$, a safe value, (see table 251) we must put $$\frac{R F_{m}}{E I} = l \quad \text{[according to eq. (2) §241.] i.e. in equation (a) above, substitute } F_{m} = [4 R l] + l c, \text{ which gives}$$ $$Gh - \frac{1}{96} \cdot \frac{R^2 E I}{E^2} = \frac{1}{4} \cdot \frac{R^2 E}{E^2} \cdot Fl - \frac{1}{4} \cdot \frac{R^2 E}{E^2} P$$ (b) having put $l = F^2 P$ ($k$ being the radius of gyration §85) and $Fl = P V$ the volume of the (prismatic) beam. From equation (b) we have the energy, $Gh$ (in ft-lbs., or inch-lbs.) of the vertical blow at the middle which the beam of Fig. 270 will safely bear, and any one unknown quantity can be computed from it, (but the mass of $G$ should not be small compared with that of the beam.) The energy of this safe impact, for two beams of the same material and similar cross-sections(similarly placed), is seen to be proportional to their volumes; while if further more their cross-sections are the same and similarly placed, the safe $Gh$ is proportional to their lengths. (These same relations hold good, approximately, beyond the elastic limits.) It will be noticed that the last statement is just the con- 314 MECHANICS OF ENGINEERING. verse of what was found in §345 for static loads, (the pressure at the centre of the beam being then equal to the weight of the safe load); for there the longer the beam and (and . . . the span) the less the safe load, in inverse ratio. As appropriate to this consideration, a quotation will be given from p. 180 of "The Standard Book of Bridges and Structures," by Sir John Anderson, London, 1894, viz. "It appears from the published experiments and statements of the Railway Commissioners, that a beam 12 feet long will only support $\frac{1}{2}$ of the load that a beam 6 feet long of the same breadth and depth will support, but that it will bear double the weight suddenly applied, as in the case of a weight falling upon it," (from the same height, should be added); "or if the same weights are used, the longer one will not break by the weight falling upon it unless it falls through twice the distance required to fracture the shorter beam." 268. Combined Flexure and Torision. Crank Shafts. Fig. 271. Let $O_1B$ be the crank, and $NO_1$ the portion projecting beyond the nearest bearing. $N \cdot P$ is the pressure of the connecting-rod against the crank-pins at a definite instant, the rotary motion being due to $P$. Let $OO_1$ be perpendicular dropped from the axis $OO_1$ of the shaft upon $P$, and $l_m$ = the distance of $P$, along the axis $OO_1$ from the cross-section $NN'$ of the shaft, close to the bearing. Let $NN'$ be a diameter of this section, and parallel to $a$. In considering the portion $NO_1B$ free, and thus exposing the circular section $l_mNN'$, we may suppose that all the stresses to be acting on its elements of this surface are the tensions (above $NN'$) and the compressions (below $NN'$) and shears $\tau$ to $NN'$, due to the bending action of $P$; and the shearing stress tan. Fig. 271. 271 FLEXURE. SPECIAL PROBLEMS. 315 gent to the circles which have $O$ as a common centre, and pass through the respective $dF's$ or elementary areas, these latter stresses being due to the twisting action of $P$. In the former set of elastic forces let $p = \frac{tensile stress per unit of area in the small parallelopiped element $m$ of the helix which is furthest from $NN'$ (the neutral axis) and $I_m$ the moment of inertia of the circle about $NN'$; then taking moments about $NN'$ for the free body, disregarding the motion) we have as in cases of flexure (see §239) $$\frac{pI}{r} = Pt ; i.e., p = \frac{Pt}{r} \quad . . . \quad (a)$$ [None of the shears has a moment about $NN'$.] Next taking moments about $OO_n$, (the flexure elastic forces, both normal and shearing, having no moments about $OO_n$) we have as in torsion (§216) $$\frac{p.I_m}{r} = Pa ; i.e., p = \frac{Pa}{I_m} \quad . . . \quad (b)$$ in which $p$, is the shearing stress per unit of area, in the torsional elastic forces, on any outermost $dF$, at as at $m$; and $I_m$ the polar moment of inertia of the circle about its centre $O$. Next consider free, in Fig. 272, a small parallelopiped taken from the helix at $m$ (of Fig. 271). The stresses [see §209] acting on the four faces $\mathbf{f}$ to the paper in Fig. 272 are there represented, the dimensions (millimetres) being $s_1$ to $NN'$, $h_1$ to $OO_n$, and $d_1$ to the paper in Fig. 272. A diagram showing a small parallelepiped cut from a helix at point \( m \), with various lines and labels indicating different forces and moments. Free. ex. P.M. H g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T g.d.f T 316 MECHANICS OF ENGINEERING. By altering the ratio of $b$ to $a$ we may make the angle $\theta$ what we please. It is now proposed to consider free the triangular prism, $GH$, to find the intensity of normal stress $q$, per unit of area, on the diagonal plane $GH$, (of length $c$) which is a bounding face of that triangular prism. See Fig. 287. By writing $x$ (perpendicu- lar to $GH$)=0, we shall have, transposing, $$q = p \sin \theta + p \cos \theta$$ and solving for $q$ $$q = p \left( \frac{b}{c} \sin \theta + \frac{a}{c} \cos \theta \right)$$ but $n = c \sin \theta$ and $b = c \cos \theta$ $$q = p \sin \theta + p \cos \theta$$ This may be written (see eqs. 63 and 60, O. W. J. Trigo- nometry) $$q = \frac{1}{2} p (1 - \cos 2\theta) + p \sin 2\theta$$ As the diagonal plane $GH$ is taken in different positions (i.e., as $\theta$ varies), this tensile stress $q$ (lbs. per sq. in. for instance) also varies, being a function of $\theta$, and its max. value may be > $p$. To find $q$ for $\max q$ we put $$\frac{dq}{d\theta} = p (\sin 2\theta + 2p \cos 2\theta)$$ -0, and obtain: $\tan [2\theta$ for $\max q] = \frac{2p}{p}$ Call this value of $\theta$, $\theta'$. Since tan $2\theta'$ is negative, $2\theta'$ lies either in the second or fourth quadrant, and hence $$\sin 2\theta' = \pm \frac{2p}{\sqrt{p^2 + 4p^2}}$$ and $$\cos 2\theta' = -\frac{p}{\sqrt{p^2 + 4p^2}}$$ [See equations 28 and 29 Trigonometry, O. W. J.] The **FLEXURE CRANK SHAFT.** Upper signs refer to the second quadrant, the lower to the fourth. If we now differentiate (4), obtaining $$\frac{dq}{ds} = 2p \cos 2\theta - 4p \sin 2\theta \quad . . . . . . . (7)$$ we note that the sine and cosine of the $[2\theta]$ of the 2nd quadrant [upper signs in (6)] are substituted in (7) the result is negative, indicating a maximum ; that is, $q$ is a maximum for $\theta = \theta'$ of eq. (6) when the upper signs are taken (2nd quadrant). To find $q$ max., then, put $\theta$ for $\theta$ in (3) substituting from (6) (upper signs). We thus find $$q_{max} = \frac{1}{2}(p + \sqrt{p^2 + 4p^2}) \quad . . . . . . . (8)$$ A similar process, taking components parallel to $GH$, Fig. 273, will yield $q$, max., i.e., the max. shear per unit of area, which for a given $p$ and $p_x$ exists on the diagonal plane $GH$ in any of its possible positions, as $\theta$ varies. This max. shearing stress is $$q_{max} = \frac{1}{2}\sqrt{p^2 + 4p^2} \quad . . . . . . . (9)$$ In the element diametrically opposite to $m$ in Fig. 271, $p$ is compression instead of tension; $q$ maximum will also be compression but is numerically the same as the $q$ max. of eq. 8. **309. Example.— In Fig. 271 suppose $P = 2$ tons = 4,000 lbs., $a = 6$ in., $l = 5$ in., and that the shaft is of wrought iron. Required its radius that the max. tension or compression shall not exceed 7,000 lbs. per sq. in.; that is, we put $q = 12,000$ in eq. (8) and solve for $r$; also $q = 7,000$ in (9) and solve for $r$. The greater value of $r$ should be taken. From equations (a) and (b) we have (see §§ 219 and 247 for $I_x$ and $I_y$) $$\frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^2y}{dx^2} = \frac{d^ 318 MECHANICS OF ENGINEERING. $$p = \frac{4Pl}{\pi r^2} \text{ and } p_s = \frac{2Pa}{\pi r^2}$$ which in (8) and (9) give $$\max.g = \frac{1}{2}\frac{P}{\pi r^2}\left[\frac{d(4l^3 + 4(2a)^3)}{dr}\right] \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8a)$$ and $$\max.q_s = \frac{1}{2}\frac{P}{\pi r^2}\sqrt{\frac{d(4l^3 + 4(2a)^3)}{dr}} \quad . . . . . . . . . . . . . . . . (9a)$$ With max. $g$=12,000, and the values of $P$, $a$, and $l$, already given, (units, inch and pound) we have from (8a), $r$=2.72 cubic inches $.\cdot$. $r$=1.39 inches. Next, with max. $q$=7,000, $P$, $a$, and $l$ as before; from (9a), $r$=1.84 cubic inches $.\cdot$. $r$=1.41 inches. The latter value of $r$=1.41 inches, should be adopted. It is here supposed that the crank-pin is in such a position (when $P$=4,000 lbs., and $q$=6 in.) that $q$ max. (and $g$, max.) are greater than for any other position; a number of trials may be necessary to decide this, since $P$ and $a$ are different with each new position of the connecting rod. If the shaft and its connections are exposed to shocks, $R$ and $S$ should be taken much smaller. 270. Another Example of combined torsion and flexure is shown in Fig. 274. The work of the working force $P$(vertical cog-pressure) is expended in overcoming the resistance (another vertical cog-pressure) $Q_r$. In the case of the rigid body consisting of the two wheels and shaft is employed to transmit power, at a uniform angular velocity, and since it is symmetrical about its axis of rotation the forces acting on it, considered free, form a balanced system. (See § 114). Hence given $P$, and the various geometrical quan- FLEXURE. CRANK SHAFT. 319 tities $a$, $b$, etc., we may obtain $Q_1$ and the reactions $P_1$ and $P_2$ in terms of $P_0$. The greatest moment of flexure in the shaft will be either $P_d$, at $C$; or $P_d$, at $D$. The portion $CD$ is under torsion, of a moment of torsion $-P_d a - Qd b$. Hence we proceed as in the example of § 309, simply putting $P_d$ (or $P_{d'}$, whichever is the greater) in place of $P_1$, and $P_a$ in place of $P_0$. We have here neglected the weight of the shaft and wheels. If $Q_1$ were an upward vertical force and hence on the same side of the shaft as $P_0$, the reactions $P_3$ and $P_4$ would be less than before, and one or both of them might be reversed in direction. 320 MECHANICS OF ENGINEERING. CHAPTER IV. FLEXURE, CONTINUED. CONTINUOUS GIRDES. 271. Definition.—A continuous girder, for present purpose, may be defined to be a loaded straight beam supported in more than two points, in which case we can no longer, as heretofore, determine the reactions at the supports from simple Statics alone, but must have recourse to the equations of the several elastic curves formed by its neutral line, which equations involve directly or indirectly the reactions sought; the latter may then be found as if they were constants of integration. Practically this amounts to saying that the reactions depend on the manner in which the beam bends; whereas in previous cases, with only two supports, the reactions were independent of the form of the elastic curves (the flexure being slight, however). As an illustration, if the straight beam of Fig. 275 is placed on three supports $O$, $B$, and $C$, at the same level, the reactions of these supports seem at first sight indeterminate; for on considering the whole beam free, we have three unknown quantities and only two equations, viz: $x$ (vertical component) = 0 and $y$ (moment about some point) = 0. If now $O$ be gradually lowered, it receives less and less pres- FLEXURE CONTINUOUS GIRDERS. 321 sure, until it finally reaches a position where the beam barely touches it; and then $O$'s reaction is zero, and $B$ and $C$ support the beam as if $O$ were not there. As to how low $O$ must sink to obtain this position, depends on the stiffness and load of the beam. Again, if $O$ be raised above the level of $B$ and $C$ receives greater and greater pressure, until the beam fails to touch one of the other supports. Still another consideration is that if the beam were torsionally rigid, its reactions would be different at $B$ and $C$, the three reactions would be different from their values for a prismatic beam. It is therefore evident that for more than two supports the values of the reactions depend on the relative heights of the supports and upon the form and elasticity of the beam, as well as upon the load. The circumstance that the beam is made continuous over the support $O$, instead of being cut apart at $O$ into two independent beams, each covering its own span and having its own reactions, shows the significance of the term "continuous girder." All the cases here considered will be comparatively simple, from the symmetry of their conditions. The beams will all be prismatic, and all external forces (i.e. loads and reactions) perpendicular to the beam and in the same plane. All supports at the same level. 272. Two Equal Spans: Two Concentrated Loads, One in the Mid- dle of Each Span. Prismatic Beam—Fig. 275. Let each half- span = $\frac{1}{2}l$. Neglect the weight of the beam. Required the reactions at $A$, through which $P_1$ acts, $P_2$ and $P_3$. From symmetry $P_2 = P_3$, and the tangent of the elastic curve at $O$ is horizontal; and since the supports are on a level the deflection of $C$ (and $B$) below $O$'s tangent is zero. The separate elastic curves $OD$ and $DC$ have a common slope and a common ordinate at $D$. For the equation of OD, make a section a anywhere between $O$ and $D$, considering $nC$ a free body. Fig. 276 (a) A diagram showing a continuous beam with two equal spans, one in the middle of each span, and two concentrated loads acting on it. 322 MECHANICS OF ENGINEERING. A diagram showing a curved beam with a neutral axis at x, and two points D and C on either side of the neutral axis. The diagram also shows a line segment from D to C, labeled as (a) and (b). 78 with origin and axis as there indicated. From $I$ (moments about neutral axis of $n$) = 0 we have (see § 231) $$E\frac{dy}{dx} = P(\frac{y_1x - x^2}{2}) - P_c(l-x)\quad \dots \quad (1)$$ $$\therefore E\frac{dy}{dx} = P(\frac{y_1x - x^2}{2}) - P_c(l-x)\quad + (C=0) \quad (3)$$ The constant $C = 0$, for at $O$ both $x$, and $dy + dx = 0$. Taking the $x$-anti-derivative of $P$ we have $$EIy = P(\frac{l^2x - x^3}{6}) - P_c\left[\frac{l^2x - x^3}{6}\right] \quad \dots \quad (3')$$ Here again the constant is zero since at $O$, $x$ and $y$ both = 0. (3) is the equation of OD, and allows no value of $x < 0$ or $\geq \frac{l}{2}$. It contains the unknown force $P_c$. For the equation of DC, let the variable section $n$ be made anywhere between $D$ and $C$, and we have (Fig. 276 b): $z$ may now range between $\frac{l}{2}$ and $l$ $$EI\frac{d^2y}{dx^2} = -P_c(l-x) \quad \dots \quad (4)$$ $$\therefore EI\frac{dy}{dx} = -P_c\left(\frac{l-z}{2}\right) + C' \quad (5')$$ To determine $C'$, put $x = \frac{l}{2}$ both in (5') and (3), and equate the results (for the two curves have a common tangent line at $D$) whence $C' = \frac{l}{4}P_c$ $$EI\frac{dy}{dx} = \frac{l}{4}P_c - P_c\left(\frac{l-z}{2}\right) \quad \dots \quad (6)$$ FLXURE. CONTINUOUS GIRDEBS. 323 Hence $EIfy = \frac{1}{2}Pp - P_0\left(\frac{w^2}{2} - \frac{p^2}{2}\right) + C'y$ . . . . (6') At $D$ the curves have the same $y$, hence put $x = \frac{1}{2}$ in the right hand member both of (3) and (6'), equating results, and we derive $C'' = -\frac{1}{2}Pp$ $\therefore EIfy = \frac{1}{2}PPx - P_0\left[\frac{w^2}{2} - \frac{p^2}{2}\right] + Pp$ . . . . (6) which is the equation of $DC$, but contains the unknown reaction $P_0$. To determine $P_0$ we employ the fact that $O's$ tangent passes through $C$, (supports on same level) and hence when $x = l$ in (6), $y$ is known to be zero. Making these substitutions in (6) we have $0 = \frac{1}{2}PP - \frac{1}{2}P_0\left[\frac{w^2}{2} - \frac{p^2}{2}\right] : P_0 = \frac{1}{4}P$ From symmetry $P_a$ also = $\frac{1}{4}P$, while $P_b$ must = $\frac{wP}{4}$, since $P_a + P_b + P_c = 2P$ (whole beam free). [Note— If the supports were not on a level, but if, (for instance) the middle support $O$ were a small distance = $h_o$ below the level line joining the others, we should put $x = l$ and $y = h_o$ in eq. (6), and thus obtain $P_a = P_b = \frac{w}{4}P + \frac{3hl_o}{8}$, which depends on the material and form of the prismatic beam and upon the length of one span, (whereas with supports all on a level, $P_a = P_b = \frac{w}{4}P$ is independent of the material and form of the beam so long as it is homogeneous and prismatic.) If $P_a$, which would then = $\frac{w}{4}P - 6EI(h_o + l)$, is found to be negative, it shows that $O$ requires a support from above, instead of below, to case it to occupy a position $h_o$ below the other supports, i.e., the beam must be "latched down" at $O$. The moment diagram of this case can now be easily constructed; Fig. 27'. For any free body $aC$, a lying in $DC$, we have 324 MECHANICS OF ENGINEERING. Diagram showing a beam with a concentrated load at point D. $$M = \frac{1}{2}Pz$$ I.e., varies directly as $z$, until $x$ passes $D$ when, for any point on $DO$, $$M = \frac{1}{2}Pz - \frac{1}{2}(x-\frac{l}{2})$$ which is 0, (point of inflection of elastic curve) for $x = \frac{l}{2}$ (note that $x$ is measured from $C$ in this figure) and at $O$, where $x = l$, becomes $\frac{1}{2}Pl$ $$\therefore M_{o} = \frac{1}{2}Pl; M_{c} = 0; M_{d} = \frac{1}{2}Pl; \text{and } M_{e} = 0$$ Hence, since $M_{max} = \frac{1}{2}Pl$, the equation for safe loading is $$\frac{RI}{e} + \frac{6}{32}Pl$$ (7) The shear at $C$ and anywhere on $CD = \frac{1}{2}P$, while on $DO$ it $- \frac{1}{2}P$ in the opposite direction. $$\therefore J_{c} = \frac{1}{2}P$$ (8) The moment and shear diagrams are easily constructed, as shown in Fig. 37, the former being symmetrical about a vertical line through $O$, the latter about the point $O^{\prime}$ Both are bounded by right lines. **273. Two Equal Spans Uniformly Distributed Load Over Whole Length of Beam.** Diagram showing two equal spans with uniformly distributed load over whole length of beam. Fig. 37a. Supports $R_1$, $R_2$, $C_1$ on a level. Total load $$= 2W = 2F$$ and may include that of the beam; $w$ is constant. As before, from symmetry $P_a = P_b$, the unknown reactions at the extremities. Diagram showing supports and reactions at ends of beams. Fig. 37b. Supports $R_1$, $R_2$, $C_1$ on a level. Total load $$= 2W = 2F$$ and may include that of the beam; $w$ is constant. As before, from symmetry $P_a = P_b$, the unknown reactions at the extremities. **FLEXURE CONTINUOUS GIRDES.** 323 Let $0 = -z$; then with $aC$ free, $x$ moves about $n=0$ gives $$EI \frac{d^2y}{dx^2} = w(l-x)\left(\frac{x}{2}-P_c(x-z)\right) - P_c\left(\frac{l-2l+x^2}{2}\right) + (Const=0) \quad (1)$$ $$\therefore EI \frac{dy}{dx} = \frac{w}{2}(lx-lx^2+\frac{x^2}{3}) - P_c\left(\frac{l-x}{2}\right) + (Const=0) \quad (2)$$ [Const.=0 for at $O$ both $dy/dx$ the slope, and $x$, are $=0$] $$\therefore EIy = \frac{w}{2}[\frac{1}{3}lx^3-\frac{1}{6}lx^2+\frac{1}{12}x^3] - P_c\left(\frac{1}{6}lx^2-\frac{1}{4}x^3\right) + (Const=0) \quad (3)$$ [Const.=0 for at $O$ both $x$ and $y$ are $=0$. Equations (1), (2), and (3) admit of any value of $x$ from 0 to $l$, i.e., hold good for any point of the elastic curve $OC$, the loading on which follows a continuous law (viz. $w=$ constant). But when $x=l$, i.e., at $C$, $y$ is known to be equal to zero, since $O$, $B$ and $C$ are on the axis of $X$, (in tangent at $O$). With these values of $x$ and $y$ in eq. (3) we have $$0 = \frac{w}{2}\cdot\frac{l}{3}\cdot\frac{1}{6}P_cP_c = y_{at}=y_{at}W$$ $$\therefore P_a=\frac{1}{6}W \text{ and } P_b=2W-2P_a=\frac{5}{6}W$$ The Moment and Shear Diagrams can now be formed since all the external forces are known. In Fig. 279 measure $x$ from $C$. Then in any section $w$ the moment of the "stress-couple" is $$M=\frac{1}{6}Wx-\frac{wx^3}{2} \quad (1)$$ which holds good for any value of $x$ on $CO$, i.e., from $x=0$ up to $l$. By inspection it is seen that for $x=0$ $$M=0; \text{ and also for } x=\frac{1}{6}l, M=0; \text{ at the inflection point } G,$$ beyond which, toward $O$, the upper fibres are in tension 326 MECHANICS OF ENGINEERING. the lower in compression, whereas between $C$ and $G$ they are vice versa. As to the greatest moment to be found on $OG$, put $dM = dx = 0$ and solve for $x$. This gives $$\frac{M}{W} = \frac{x}{L} : [x \text{ for } M_{\text{max}}] = \frac{3L}{4} \quad (2)$$ which in eq. (1) gives $$M_x (\text{at } N, \text{ see figure}) = \frac{3L}{4} W \quad (2)$$ But this is numerically less than $M_d = -\frac{3L}{4} W$! Hence the stress in the outer fibre at $O$ being $$p_o = \frac{3L}{4} \frac{W}{I} \quad (3)$$ the equation for safe loading is $$\frac{EI}{c} = \frac{3L}{4} W \quad (4)$$ the same as if the beam were cut through at $O$, each half, of length $l$, retaining the same load as before (see § 243 eq. (2)). Hence making the girder continuous over the middle support does not make it any stronger under a uniformly distributed load; but it does make it considerably stiffer. As for the shear, $J$, we obtain it for any section by taking the $x$-derivative of $M$ in eq. (1), or by putting 2(vertical forces) = 0 for the free body at $C$, and thus have for any section on $OO$ $$J = \frac{3L}{4} W - wx \quad (5)$$ $J$ is zero for $x = \frac{3L}{4}$ (where $M$ reaches its calculus maximum $M_{\text{max}}$; see above) and for $x = L$ it = -$\frac{3L}{4} W$ which is numerically greater than $\frac{3L}{4} W$, its value at $C$. Hence $$J_c = \frac{3L}{4} W \quad (6)$$ FLEXURE. CONTINUOUS GIRDERS. 327 The moment curve is a parabola (a separate one for each span), the shear curve a straight line, inclined to the horizontal, for each span. **Problem.** How would the reactions in Fig. 278 be changed if the support $O$ were lowered a (small) distance $h$, below the level of the other two? 274. *Prismatic Beam Fixed Horizontally at Both Ends (at Same Level). Single Load at Middle.* Fig. 280. [As usual the beam is understood to be homogeneous so that $E$ is the same at all sections]. The building in, or fixing, of the two ends is supposed to be of such a nature as to cause no horizontal constraint; i.e., the beam does not act as a cord or chain, in any manner, and hence the sum of the horizontal components of all forces acting on any section is zero, as in all preceding cases of flexure. In other words the neutral axis still contains the centre of gravity of the section and the tensions and compressions are equivalent to a couple (the stress-couple) whose moment is the "moment of flexure." If the beam is conceived cut through close to both wall faces, and this portion of length $l$, considered free, the forces holding it in equilibrium consist of the downward force $P$ (acting at the middle point of each section); and two "stress-couples" - one in each section, whose moments are $M_1$ and $M_2$. From symmetry we know that $J_{1} = J_{2}$ and that $M_1 = M_2$. From $Y = 0$ for the free body just mentioned, (but not shown in the figure), and from symmetry, we have $J_{1} = \frac{1}{2} P$ and $J_{2} = \frac{1}{2} P$; but to determine $M_1$ and $M_2$, the form of the elastic curves $OB$ and $BO$ must be taken into account as follows: *Equation of OB.* Fig. 280. $\Sigma$ (mom. about neutral axis of any section) on $OB$ = 0 (for the free body w/o which) A diagram showing a prismatic beam fixed horizontally at both ends (at same level). A single load is placed at the middle. The beam is assumed to be homogeneous so that E is the same at all sections. The building in, or fixing, of the two ends is supposed to be of such a nature as to cause no horizontal constraint. 328 MECHANICS OF ENGINEERING. has a section exposed at each end, $a$ being the variable section) will give $$E \frac{dF}{dx} = P(\frac{1}{2} - x) + M - \frac{1}{2} P(1 - x) \quad . . . (1)$$ (Note. In forming this moment equation, notice that $M$ is the sum of the moments of the tensions and compressions at $C$ about the neutral axis at $n$, just as much as about the neutral axis of $C$; for those tensions and compressions are equivalent to a couple, and hence the sum of their moments is the same taken about any axis whatever $\gamma$ to the plane of the couple (§39). Taking the $x$-anti-derivative of each member of (1), $$E \frac{dy}{dx} = P(\frac{1}{2} l - x) + M - \frac{1}{2} P(l - x) \quad . . . (2)$$ (The constant is not expressed, as it is zero). Now from symmetry we know that the tangent-line to the curve $OB$ at $B$ is horizontal, i.e., for $x = \frac{1}{2} l$, $dy/dx = 0$, and these values in eq. (2) give $$0 = \frac{1}{2} P\left(\frac{1}{2} l\right) + M - \frac{1}{2} P\left(\frac{1}{2} l\right); \text{ hence } M = M = \frac{1}{2} P\left(\frac{1}{2} l\right) \quad . . . (3)$$ Safe Loading. Fig. 251. Having now all the forces which act as external forces in straining the beam $OC$, we are ready to draw the moment diagram and find $M$. For convenience measure $\sigma$ from $C$. For the free body at $c$, we have [see eq. (3)] $$P\sigma - M + \frac{\pi F}{e} = 0; \text{ hence } M = \frac{1}{2} P\left(\frac{1}{2} l\right) \quad . . . (4)$$ Eq. (4) holds good for any section on $CB$. By putting $x = 0$ we have $M = M = \frac{1}{2} P\left(\frac{1}{2} l\right)$; lay off $H C' = M$ to scale (so many inch-pounds means so many pounds per foot per inch), At $B$, for $x = \frac{1}{2} l$, $M = \frac{1}{2} P\left(\frac{1}{2} l\right)$; hence lay off $H B' = \frac{1}{2} P\left(\frac{1}{2} l\right)$ on the opposite side of the axis $OC'$. Fig. 31. 3 FLEXURE CONTINUOUS GIRDERS. 329 from $HC'$, and join $DH$. $DK$, symmetrical with $DH$ about $BD$, completes the moment curves, viz.: two right lines. The max. $M$ is evidently = $\frac{1}{4}PL$ and the equation of safe loading $$\frac{RI}{e}=\frac{1}{4}PL \quad \quad \quad \quad (5)$$ Hence the beam is twice as strong as if simply supported at the ends, under this load ; it may also be proved to be four times as stiff. The points of inflection of the elastic curve are in the middles of the half-spans, while the max. shear is $$J_n=\frac{1}{2}P \quad \quad \quad \quad (6)$$ 275. Pneumatic Beam Fixed Horizontally at Both Ends (at Same Level). Uniformly Distributed Load Over the Whole Length. Fig. 282. As in the preceding problem, we know from symmetry that $J_n=J_m=\frac{1}{2}W=\frac{1}{2}wt$, and that $M_0=M_w$, and determine the latter quantities by the equation of the curve $OC$, there being but one curve in the present instance, instead of two, as there is no change in the law of loading between $O$ and $C$. With a $O$ free, $\frac{d^2y}{dx^2}=0$ gives $$EI\frac{d^2y}{dx^2}=-\frac{1}{4}Wx+M_0+\frac{wx^2}{2} \quad \quad \quad (1)$$ and . . . $$EI\frac{d^2y}{dx^2}=-\frac{1}{4}W\frac{x^2}{2}+M_w+\frac{wx^2}{2}+(C=0) \quad \quad (2)$$ Diagram showing a flexure girder with a uniformly distributed load over its entire length. Fig. 282. 330 MECHANICS OF ENGINEERING. The tangent line at $O$ being horizontal we have for $x=0$, $\frac{dy}{dx}=0, \ldots C=0$. But since the tangent line at $C$ is also horizontal, we may for $x=l$ put $dy+dz=0$, and obtain $$0=-\frac{1}{4}Wl+MJ+\frac{1}{4}wx^2,$$ whence $M_c=\frac{1}{4}Wl \quad (3)$$ as the moment of the stress-couple close to the wall at $O$ and at $C$. Hence, Fig. 283, the equation of the moment curve (a single continuous curve in this case) is found by putting $\Sigma(mom.)=0$ for the free body $NO$, of length $x$, thus obtaining Fig. 283. $$\frac{P}{e}+\frac{1}{2}Wx-M_c-\frac{wx^2}{2}=0$$ I.e., $$M=\frac{1}{4}Wl+\frac{wx^2}{2}-\frac{wx^2}{2} \quad (4)$$ an equation of the second degree, indicating a conic. At $O$, $M=M_c=\frac{1}{4}Wl$; at $B$ by putting $x=\frac{1}{2}l$ in (4), we have $M_c=-\frac{1}{4}Wl$, which is less than $M_c$, although $M_c$ is the calculus max. (negative) for $M_c$ as may be shown by writing the expression for the shear $(J-\frac{1}{4}W-wx)$ equal to zero, etc. FLEXURE. CONTINUOUS GIRDERS. 331 Hence $M_{n} = \frac{1}{2} WL$, and the equation for safe loading is $$\frac{R' I}{c} = \frac{1}{2} WL \dots \dots \dots (5)$$ Since (with this form of loading) if the beam were not built in but simply rested on two end supports, the equation for safe loading would be $(RI - e) = \frac{1}{4} WL$, (see §242), it is evident that with the present mode of support it is 50 per cent. stronger as compared with the other; i.e., as regards normal stresses in the outer elements. As regards shearing stresses in the web if it has one, it is no stronger, since $J_{w} = \frac{1}{2} WL$ in both cases. As to significant points of the uniform load, the max. deflection in the present case may be shown to be only $\frac{1}{8}$ of that in the case of the simple end supports. It is noteworthy that the shear diagram in Fig. 283 is identical with that for simple end supports §243, under uniform load; while the moment diagrams differ as follows: The parabola $K'P'd$, Fig. 283, is identical with that in Fig. 235, but the horizontal axis from which the ordinates of the former are measured, instead of joining the extremities of the curve, cuts it in such a way as to have equal areas between it and the curve, on opposite sides. i.e., area's $[KC'P' + AG'G']$ = area $H'G'B'$ In other words, the effect of shifting the ends horizontally is to shift the moment parabola upward a distance $- M_n$ (to scale), $\frac{1}{2} WL$, with regard to the axis of reference, $O'B'$, in Fig. 283. **Remarks.—The foregoing very simple cases of continuous girders illustrate the means employed for determining the reactions of supports and eventually the max. moment and the equations for safe loading and for deflections. When there are more than three supports, with spans of unequal length, and loading of any description, the analysis leading to the above results is much more complicated and tedious, but is considerably simplified** A diagram showing a continuous girder with various loads applied. 332 MECHANICS OF ENGINEERING. and systematized by the use of the remarkable theorem of three moments, the discovery of Clapeyron, in 1857. By this theorem, given the spans, the loading, and the vertical heights of the supports, we are enabled to write out a relation between the moments of each three consecutive supports, and thus obtain a sufficient number of equations to determine the moments at all the supports [p. 641 Rankine's Applied Mechanics.] From these moments the shears close to any section of the supports are found, then the reactions, and from these and the given loads the moment at any section can be determined; and hence finally the max. moment $M_{\text{max}}$ and the max. shear $J_{\text{max}}$. The treatment of the general case of continuous girders by graphic methods, however, is comparatively simple, and its presentation is therefore deferred, § 391. THE DANGEROUS SECTION OF NON-PRIS- MATIC BEAMS. 277. Remark. By "dangerous section" is meant that section (in a given beam under given loading with given mode of support) where $p$, the normal stress in the outer fibre, at distance $\epsilon$ from its neutral axis, is greater than in the outer fibre of any other section. Hence the elasticity of the material will be first impaired in the outer fibre of this section, and will gradually increased in amount (but not altered in distribution). In all preceding problems, the beam being prismatic, $I$, the moment of inertia, and $\epsilon$ were the same in all sections, hence when the equation $\frac{F}{e} = M$ [§389] was solved for $p$, giving $$p = \frac{Mc}{I} \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1)$$ we found that $p$ was a max., $= p_{\text{max}}$, for that section whose $M$ was a maximum, since $p$ varied as $M$, for the moment **PLURURE NON-PRISMATIC BEAMS.** 333 of the stress-couple, as successive sections along the beam were examined. But for a non-prismatic beam $I$ and $c$ change, from section to section, as well as $M$, and the ordinate of the moment diagram no longer shows the variation of $p$, nor is $p$ a max. where $M$ is a max. To find the dangerous section, then, for a non-prismatic beam, we express the $M$, the $I$, and the $c$ of any section in terms of $x$, thus obtaining $p = \text{func.}(x)$, then writing $dp + dx = 0$, and solving for $x$. 278. Dangerous Section in a Double Truncated Wedge. Two End Supports. Single Load in Middle. The form is shown in Fig. 284. Neglect weight of beam ; measure $z$ from one support $O$. The reaction at each support is $\frac{1}{2}P$. The width of the beam $b$ at all sections, while its height, $u$, varies, being $h$ at $O$. To express the $e = \frac{1}{2}v$, and the $I = \frac{1}{3}bu^3 (\S 247)$ of any section on $OC$, in terms of $x$, conceive the sloping faces of the truncated wedge to be prolonged to their intersection $A$, at a known distance $c$ from the face at $O$. We then have from similar triangles $$v : x + c : h : e, v = \frac{a}{b}(x + c) \quad . . . (1)$$ and $$e = \frac{1}{2}\frac{h}{c}(x + c) \text{and } I = \frac{1}{3}\frac{h^3}{c}(x + c)^3 \quad . . . (2)$$ For the free body $\neq O$, $\Sigma$ (moments) = 0 gives $$\frac{1}{2}Pz - \frac{p}{e}f = 0 : p - Pze \quad . . . (3)$$ (That is, the $M = \frac{1}{4}Fx$. But from (2), (3) becomes) $$p = 3P\frac{d}{dx}\cdot\frac{x}{(x+c)}; \text{ and } dp = 3P\frac{d}{dx}\cdot\frac{(x+c)-2(x+c)}{(x+c)^2} \quad (4)$$ By putting $dp + dx = 0$ we obtain both $x = -a$, and 334 MECHANICS OF ENGINEERING. $x + c$, of which the latter, $x = +c$, corresponds to a maximum for $p$ (since it will be found to give a negative result on substitution in $\frac{dp}{dx} + dx^2$). Hence the dangerous section is as far from the support $O$, as the imaginary edge, $A$, of the completed wedge, but of course on the opposite side. This supposes that the half-span, $\frac{1}{2}l$, is > $c$; if not, the dangerous section will be at the middle of the beam, as if the beam were prismatic. Hence, with $\frac{1}{2}l < c$ the equation for safe loading is ($h$=height) $\frac{Rbh^2}{6} = \frac{1}{2}Pl$ (5) while with $\frac{1}{2}l > c$ and $p = R'$ in (3) (see §239.) $\frac{1}{2}l < c$ the equation for safe loading is (put $x = \frac{1}{2}l$) $\frac{Rb(2h)^2}{6} = \frac{1}{2}Pc$ (6) \begin{align*} &\text{(see §239.)}\\ &\text{279. Double Truncated Pyramid and Cone. Fig. 285. For}\\ &\text{the truncated pyramid both width = } u,\text{ and height = } v,\text{ are variable, and if } b\text{ and } h\text{ are the dimensions at } O,\text{ and }\\ &c = OA\text{ = distance from } O\text{ to the imaginary vertex } A,\text{ we}\\ &\text{shall have from similar triangles } u = \frac{b}{c}(x+c)\text{ and } v = \frac{h}{c}(x+c).\text{ Hence, substituting } e = \frac{1}{2}v\text{ and } I = \frac{1}{3}u^3,\text{ in the moment equation}\\ &\frac{pI}{e} - \frac{P_2}{2} = 0,\text{ we have } p = 3P\left(\frac{x}{b}\right)^4\left(\frac{x+c}{x+c}\right)^{-3}\cdot\cdot\cdot (7)\\ &\therefore \frac{dp}{dx} = 3P\left(\frac{x}{b}\right)^4\left(\frac{x+c}{x+c}\right)^{-3}\cdot\cdot\cdot (8) \end{align*} Fig. 285. Fig. 285. FLEXURE NON-PRISMATIC BEAMS. 335 Putting this - 0, we have $x = -c$, $z = -c$, and $x + \frac{1}{2} c$, $e$ being the dangerous section at a distance $x = \frac{1}{2} c$ from $O$, and the equation for safe loading is either $\frac{Rb^2l^2}{6} - \frac{1}{4} Pl \ldots$ if $\frac{1}{2} l$ is less than $\frac{1}{2} c$ . . . . (9) (in which $b'$ and $h'$ are the dimensions at mid-span) or $\frac{R(1/2 b)(\frac{1}{2} c h')^2}{6} = \frac{1}{4} Pc$ if $\frac{1}{2} l$ is greater than $\frac{1}{2} c$ . . . . (10) For the truncated case (see Fig. 385 also, on right) where $e$ = the variable radius $r$, and $I = \frac{1}{4} \pi r^4$, we also have $p = [Const.] \cdot \frac{x}{(x+c)^2}$ . . . . (11) and hence $p$ is a max. for $x = \frac{1}{2} c$, and the equation for safe loading either $\frac{\pi Rr^2a}{4} = \frac{1}{4} Pl$, for $\frac{1}{2} l < \frac{1}{2} e$ . . . . (12) (where $r' =$ radius of mid-span section); or $\frac{\pi R^2(\frac{1}{2} r_0)^2}{4} = \frac{1}{4} Pe$, for $\frac{1}{2} l > \frac{1}{2} e$ . . . . (13) (where $r_0 =$ radius of extremity.) NON-PRISMATIC BEAMS OF "UNIFORM STRENGTH." 380. Remarks A beam is said to be of "uniform strength" when its form, its mode of support, and the distribution of loading are such that the normal stress has the same value in all the outer fibres, and thus one element of economy is secured, viz.: that all the outer fibres may be made to do full duty, since under the safe loading, $p$ will be equal to $E$ in all of them. [Of course, in all cases of flexure, the elements between the neutral surface and 336 MECHANICS OF ENGINEERING. the outer fibres being under tensions and compressions less than $H$ per sq. inch, are not doing full duty, as regards economy of material, unless perhaps with respect to shearing stresses.) In Fig. 265, §201, we have already had an instance of a body of uniform strength in flexure, viz.: the middle segment, $CD$, of that figure; for the moment is the same for all sections of $CD$ (eq. (2) of that §), and hence the normal stress $p$ in the outer fibres (the beam being prismatic) is that which would exist if the weight of the beam itself is neglected. The general method pursued will be to find an expression for the outer-fibre-stress $p$, at a definite section of the beam, where the dimensions of the section are known or assumed, then an expression for $p$ in the variable section, and equate the two. For clearness the figures are exaggerated, vertically. 281. Parabolic Working Beam. Unsymmetrical. Fig. 286 A diagram showing a parabolic working beam with various sections labeled A, B, C, D, E, F. Fig. 286 CBO is a working beam or lever, $B$ being the fixed fulcrum or bearing. The forces $P$, being given we may compute $P_c$ from the mom equation $P_dL = P_cL$, while the fulcrum reaction is $P_a = P_b + P_c$. All the forces are $\mathbf{T}$ to the beam. The beam is to have the same width $b$ at all points, and is to be rectangular in section. Required first, the proper height $h_0$, at $B$, for safety. From the free body $BO$, of length $L$, we have $\Sigma(mom_a) = 0$; i.e., $$\frac{P_bL}{c} = P_cL; \text{or } p_b = \frac{6PL}{bh_0} \quad \ldots \quad (1)$$ FLEXURE NOX-PRISMATIC BEAMS. 337 Hence, putting $p_0 = E^{'}, h$, becomes known from (1). Required, secondly, the relation between the variable height $v$ (at any section $n$) and the distance $x$ of $a$ from $O$. For the free body $nO$, we have ($\Sigma$ moma, = 0) $$\frac{p_0}{e} = P_x; \text{ or } \frac{P_0}{h} \cdot \frac{be^{'}}{y_2 e'} = P_x \text{ and } p_0 = \frac{6P_x}{be'}$$ (2) But for "uniform strength" $p_0$ must be $p_x$; hence equate their values from (1) and (2) and we have $$x = \frac{l_1}{h^{'}}$$ which may be written ($\frac{1}{2} y^{'})^2 = (\frac{1}{2} h^{'})^2 x$$ (3) so as to make the relation between the abscissa $x$ and the ordinate $\frac{1}{2} y'$ more marked; it is the equation of a parabola, whose vertex is at $O$. The local stresses at $C$, $B$, and $O$ must be properly provided for by evident means. The shear $J = P_n$ at $O$, also requires special attention. This shape of beam is often adopted in practice for the working beams of engines, etc. The parabolic outlines just found may be replaced by trapezoidal forms, Fig. 287, without using much more material, and by making the slope-plunging plane face parallel to the base of the trapezoid, so that the parabolic outline at points $T_1$ and $T_2$, half-way between $O$ and $B$, and $C$ and $B$, respectively. It can be proved that they contain minimum volume, among all trapezoidal forms capable of circumscribing the given parabolic bodies. The dangerous sections of these trapezoidal bodies are at the tangent points $T_1$ and $T_2$. This is as it should be, (see § 278), remembering that the subtangent of a parabola is bisected by the vertex. 282. I-Beam of Uniform Strength—Support and load same as in the preceding §. Fig. 288. Let the area of the A diagram showing a cross-section of an I-beam with a trapezoidal cross-section. Fig. 287 338 MECHANICS OF ENGINEERING. A diagram showing a flange-section with a vertical line through its center, labeled "x". The diagram also shows a horizontal line at the top and bottom of the flange-section, labeled "P". The diagram is labeled "Fig. 288" on the right side. flange-sections be $F$ and let it be the same for all values of $x$. Considering all points of $F$ at any one section as at the same distance $z$ from the neutral axis, we may write $I = x^2 F$, and assuming that the flanges take all the tension and compression while the (thin) web carries the shear, the free body of length $x$ in Fig. 288 gives (moments about $n$) $$\frac{pI}{e} = P_x; \text{i.e. } \frac{px^2F}{z} = P_x; \text{or, since } p \text{ is to be constant,}$$ $$z = [\text{Const.}], x$$ (1) i. e. $x$ must be made proportional to $z$. Hence the flanges should be made straight. Practically, if they unite at $C$, the web takes but little shear. 283. Rectang. Section. Height Constant. Two Supports (at Extremities). Single Eccentric Load. - Fig. 289. $b$ and $h$ are the dimensions of the section at $B$. With $BO$ free we have $$\frac{p_0L_o}{e_b} - P_{o0} = 0; \therefore p_x = \frac{6P_{o0}}{bh^2} \quad (1)$$ At any other section on $BO$, as $n$, where the width is $u$, the variable whose relation to $x$ is required, we have for $nO$ free $$\frac{p_nL_n}{e_n} - P_{n0}; \text{or } p_n/u^2h^2 = P_{n0}; \therefore p_n = \frac{6P_{n0}}{uh^2}$$ (2) Equating $p_n$ and $p_x$ we have $\therefore z : z_o$ (3) A diagram showing a rectangular section with a vertical line through its center, labeled "x". The diagram also shows a horizontal line at the top and bottom of the section, labeled "P". The diagram is labeled "Fig. 289" on the right side. FLEXURE BEAMS OF UNIFORM STRENGTH. 339 That is, $BO$ must be wedge-shaped with its edge at $O$, vertical. 284. Similar Rectang. Sections Otherwise as Before.—Fig. 389 a, $b$ and $h$ are the dimensions at $B$; at any other section $n$, on $BO$, the height $v$ and width $u$, are the variables whose relation to $x$ is desired and by hypothesis are connected by the relation $$w : v : b : h \quad \quad (1)$$ (since the section at $n$ is a rectangle similar to that at $B$). For the free body $BO \ldots p_a = \frac{6P_{ac}}{bh} \quad \quad (2)$ For the free body $nO \ldots p_b = \frac{6P_{ac}}{wu} \quad \quad (3)$ Writing $p_a = p_b$ we obtain $l + bh = z + wu$, in which put $u = bv + h$, from (1); whence $$\frac{v^3}{b^3} = x; \text{or} (\frac{1}{2}v)^2 = (\frac{1}{2}h)^2 \quad \quad (4)$$ which is the equation of the curve (a cubic parabola) whose abscissa is $x$ and ordinate $\frac{1}{2}v$; i.e., of the upper curve of the outline of the central longitudinal vertical plane section of the body (dotted line $BO$) which is supposed symmetrical about such a plane. Similarly the central horizontal plane section will cut out a curve a quarter of which (dotted line $Z'O'$) has an equation $$(\frac{1}{2}y)^2 = (\frac{1}{2}h')^2 \quad \quad (5)$$ That is, the height and width must vary as the cube root 340 MECHANICS OF ENGINEERING. of the distance from the support. The portion $CB$ will give corresponding results, referred to the support $C$. [If the beam in this problem is to have circular cross-sections, let the student treat it in the same manner.] 286. Uniform Load. Two End Supports. Rectangular Cross-Sections. Width Constant. Weight of beam neglected. How should the height vary, at the middle being $a$ and $b$? Fig. 289 b. From symmetry each reaction is $\frac{1}{2}W = \frac{1}{2}wL$. At any cross section $x$, the width is $= b_x$ (same as that at the middle) and the height is $= v_x$ variable. $I_x(moms_a) = 0$, for the free body $BO$, gives $$\frac{p_{x}}{e_x} = \frac{wx}{2}; i.e., \frac{p_{x}}{e_x}dx = \frac{wx}{2}dx$$ $$\frac{wx^2}{2} \cdot \ldots \quad (1)$$ But for a beam of uniform strength, $p_x$ is to be $= p_a$ as computed from $I_x(moms_a) = 0$ for the free body $.. BO$, i.e from $$\frac{p_{x}}{e_x}dx = \frac{wx}{2}dx$$ $$\frac{wx^2}{2} \cdot \ldots \quad (2)$$ Hence solve (1) for $p_a$ and (2) for $p_a$ and equate the results, whence $v^2 = \frac{h}{(y/x)^2}(L-x)^2$; or $(y/x)^2 = \frac{(y/h)^2(L-x)^2}{(y/x)^2}$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3) This relation between the abscissa $x$ and the ordinate $\frac{1}{2}v_x$ of the curve $CBO$, shows it to be an ellipse since eq. (3) is that of an ellipse referred to its principal diameter and the tangent at its vertex as co-ordinate axes. In this case eq. (3) covers the whole extent of both upper and lower curves, i.e. the complete outline, of the curve $CBOB'$, whereas in Figs. 286, 287, and 289a, such is not the case. FLEXURE BODIES OF UNIFORM STRENGTH 341 287. Cantilevers of Uniform Strength.—Beams built in at one end, horizontally, and projecting from the wall without support at the other, should have the forms given below, for the given cases of loading, if all cross-sections are to be rectangular and the weight of beam neglected. Sides of sections horizontal and vertical. Also, the sections are symmetrical about the axis of the piece. $b$ and $h$ are the dimensions at the wall. $l$ = length. No proofs given. (a) (b) (c) Fig. 290. (d) (e) (f) Fig. 291. Width constant. Vertical outline parabolic. Single end load. Height constant. Single end load. Horizontal outline triangular. Constant ratio of height $v$ to width $w$. Both outlines cubic parabolas. $$\left(\frac{3}{4}v\right)^{\frac{2}{3}}=\left(\frac{3}{4}h\right)^{\frac{2}{3}} \quad (1)$$ $$\left(\frac{3}{4}u\right)=\left(\frac{3}{4}b\right)^{\frac{2}{3}} \quad (2)$$ $$\left(\frac{3}{4}v\right)^{\frac{2}{3}}=\left(\frac{3}{4}h\right)^{\frac{2}{3}} \quad (3)$$ 342 MECHANICS OF ENGINEERING. Uniform Load. Width constant. Vertical outline triangular. Fig. 291, (a). $(\frac{1}{2}w) = (\frac{1}{2}h)\frac{z}{l}$ . (4) Uniform Load. Height constant. Horiz. outline is two parabolas meeting at $O$ (vertex) with geomet axes I to wall. Fig. 291, (b). $\frac{1}{2}w = (\frac{1}{2}b)\frac{x^2}{F}$ . (5) Uniform Load. Both outlines semi-cubic parabola. Sections similar rectangles. Fig. 291, (c). $(\frac{1}{2}w)^{\prime} = (\frac{1}{2}b)^{\prime}\frac{x^2}{F}$ . (6) 290—Beams and castilevers of circular cross-sections may be dealt with similarly, so that the proper longitudinal outline given will be such that they "body" uniform strength." As a consequence of the possession of this property, with loading and mode of support of specific character, the following may be stated; that to find the equation of safe loading any cross-section whatever may be employed. This refers to tension and compression. As regards the shearing stresses in different parts of the beam the condition of "uniform strength" is not necessarily obtained at the same time with that for normal stress in the outer fibres. DEFLECTION OF BEAMS OF UNIFORM STRENGTH. 290. Case of § 283, the double wedge, but symmetrical, i.e., $l_1 = l_2 = \frac{1}{2}l$, Fig. 292. Here we shall find the use of the BEAMS OF UNIFORM STRENGTHIL 343 (a) Diagram showing a beam with a uniform cross-section and a concentrated load at one end. Fig. 292 form $EI$ (of the three forms for the moment of the stress couple, see eqs. (5), (6) and (7), §§ 299 and 231) of the most direct service in determining the form of the elastic curve $OB$, which is symmetrical, and has a common tangent at $B$, with the curve $BC$. First to find the radius of curvature, $\rho$, at any section $n$, we have for the free body $mO_1$, $$\Sigma(mo_{n}=0), \text{ whence}$$ $$-\frac{EI}{P}xPz=0; \text{ but } (3) \S 283 \quad z=\frac{u}{b}l \text{ and } l=\frac{v}{h}w^2$$ we have $\frac{1}{\rho}\frac{E}{P}u^2=\frac{1}{b}P\frac{ul}{b}$ and $\rho=\frac{1}{\frac{u^2}{b}}\frac{bb^3}{T}\frac{E}{P}$ (1) from which all variables have disappeared in the right hand member; i.e., $\rho$ is constant, the same at all points of the elastic curve, hence the latter is the arc of a circle, having $b$ as its tangent at $B$. To find the deflection, $d$, at $B$, consider Fig. 292, (8) where $d=KH$, and the full circle of radius $BH-w$ is drawn. The triangle $KOB$ is similar to $YOB$, and : $KB:OB::OB:YB$ But $OB=\frac{y_0}{h}$, $KB=b$ and $YB=2p$ $\therefore d=\frac{(y_0)^2p}{2p}$ and $\therefore$, from eq. (1), $d=\frac{y_0^2p}{bh^2E}$ (2) From eq. (4) § 233 we note that for a beam of the same material but *prismatic* (parallelepipedical in this case,) having the same dimensions, $b$ and $h$, at all sections as at the middle, deflects an amount $\frac{1}{48}\frac{PE}{EI}=\frac{y_0^2p}{bh^2E}$ under a 344 MECHANICS OF ENGINEERING. load $P$ in the middle of the span. Hence the tapering beam of the present § has only $\frac{1}{2}$ the stiffness of the prismatic beam, for the same $b$, $h$, $l$, $E$, and $P$. 281. Case of § 281 (Parabolic Body). With $l = l_0$, i.e., Symmetrical—Fig. 293(a). Required the equation of the neutral line $OB$. For the free body $nO$, $(\text{mom}_n) = 0$ gives us $$\frac{d^2y}{dx^2} - \frac{1}{2}Pz \quad \quad \quad (1')$$ Fig. 293(b), shows simply the geometrical relations of the problem, position of origin, axes, etc. On $OB$ is the neutral line or elastic curve whose equation, and greatest ordinate $d$, are required. (The right hand member of eq. (1') is made negative because $dy/dx$ is negative, the curve being concave to the axis $X$ in this, the first quadrant.) Now if the beam were prismatic, i.e., the "moment of inertia" of the cross-section would be constant, i.e., the same for all sections, we might proceed by taking the x-anti-derivative of each member of (1') and add a constant; but it is variable and in $$= \frac{1}{2} b h^2 = \frac{1}{2} \left( \frac{bh^2}{l} \right)_T z^2 \quad \quad (\text{from eq. } 3, \text{ § 281}, \text{ putting } L = \frac{1}{2}l)$$ hence (1") becomes $$\frac{1}{2} E \left( \frac{bh^2}{l} \right)_T z^2 \frac{d^2y}{dx^2} = -\frac{1}{4} Pz \quad \quad \quad (1'')$$ To put this into the form Const. × $\frac{d^2y}{dx^2}$ = fune. of $(z)$, we need BEAMS OF UNIFORM STRENGTH. 345 only divide through by $x^{\frac{1}{2}}$, (and for brevity denote $$\frac{1}{2}Ebh_1 + (\frac{1}{2})h_1^{\frac{1}{2}}$$ by A) and obtain $$A \frac{dy}{dx} = -\frac{1}{2}P x^{-\frac{1}{2}} \quad \ldots \quad (1)'$$ We can now take the $x$-anti-derivative of each member, and have $$A \frac{dy}{dx} = -\frac{1}{2}P(2x + \frac{1}{2}) + C \quad \ldots \quad (2)'$$ To determine the constant $C$, we utilize the fact that at $B$, where $x=\frac{1}{2}$, the slope $\frac{dy}{dx}$ is zero, since the tangent line is there horizontal, whence from $(2)'$ $$0 = P(\sqrt{\frac{1}{2}} + \frac{1}{2}) + C = P(\sqrt{\frac{1}{2}} - \frac{1}{2}) \quad \ldots \quad (3)$$ :. $(2)'$ becomes $A \frac{dy}{dx} = P(\sqrt{\frac{1}{2}} - x - \frac{1}{2}) + [C' = 0] \quad \ldots \quad (3)$ :. $Ay = P[\sqrt{\frac{1}{2}}x - \frac{1}{2}x^{\frac{1}{2}}] + [C' = 0] \quad \ldots \quad (3)$ $(C' = 0$ since for $x=0, y=0).$ We may now find the deflection $d$ (Fig. 293(b)) by writing $y = \frac{1}{2}x$ and $y = d$, whence, after restoring the value of the constant $A$, $$d = \frac{1}{2}\frac{Pd}{Ebh_0} \quad \ldots \quad (4)$$ and is twice as great [being $= 2$. $\frac{PP'}{4Eb_{h_0}}$] as if the girder *See 388, putting $I = \frac{1}{3}bh_0^3$ in eq. (6). were parallelopipedical. In other words, the present girder is only half as stiff as the prismatic one. **293. Special Problem. (L.)** The symmetrical beam in Fig. 294 is of rectangular cross-section and constant width $= b$, A diagram showing a symmetrical beam with a rectangular cross-section and constant width. 346 MECHANICS OF ENGINEERING. but the height is constant only over the extreme quarter spans, being $h = \frac{1}{2}h$, i. e., half the height $\lambda$ at mid-span. The convergence of the two truncated wedges forming the middle quarters of the beam is such that the prolongations Diagram showing a beam with truncated wedges converging to form the middle quarters. of the upper and lower surfaces would meet over the supports (as should be the case to make $h=2h$). Neglecting the weight of the beam, and placing a single load in middle, it is required to find the equation for safe loading; also the equations of the four elastic curves; and finally the deflection. The solutions of this and the following problem are left to the student, as exercises. Of course the beam here given is not one of uniform strength. 293. Special Problem. (II). Fig. 295. Required the manner in which the width of the beam must vary, the height being constant, cross-sections rectangular, weight of beam neglected, to be a beam of uniform strength, if the load is uniformly distributed? Diagram showing a beam with varying widths and uniform cross-sections. FLEXURE OBLIQUE FORCES 347 CHAPTER V. FLEXURE OF PRISMATIC BEAMS UNDER OBLIQUE FORCES. 994. Remarks By "oblique forces" will be understood external forces not perpendicular to the beam, but those external forces will be confined to one plane, called the force-plane, which contains the axis of the beam and also cuts the beam symmetrically. The curvature induced by these external forces will as before be considered very slight, so that distances measured along the beam will be treated as unchanged by the flexure. It will be remembered that in previous problems the proof that the neutral axis of each cross section passes through the centre of gravity was based on the fact that when a portion of the beam having a given section as one of its bounding surfaces is considered free, the condition of equilibrium $\Sigma$ (compms. $\parallel$ to beam) = 0 does not introduce any of the external forces, since these in the problems referred to, were $\parallel$ to the beam; but in the problems of the present chapter such is not the case, and hence the neutral axis does not necessarily pass through the centre of gravity of any section, and in fact may have only an ideal, geometrical existence, being sometimes entirely outside of the section; in other words, the fibers whose ends are placed in a given plane may be either tensile or all in compression, of intensities varying with the distance of each from the neutral axis. It is much more convenient, however, to take for an axis of moments the gravity axis parallel to the A page from a book with text discussing flexure of prismatic beams under oblique forces. 348 MECHANICS OF ENGINEERING. neutral axis instead of the neutral axis itself, since this gravity axis has always a known position. 295. Classification of the Elastic Forces, Shear, Thrust, and Stress-Coins. Fig. 296. Let $AKM$ be one extremity of a portion, considered free, of a prismatic beam, under oblique forces. $C$ is the centre of gravity of the section exposed, and $GC$ the gravity axis $\mathfrak{q}$ to the force plane $JAK$. The stresses acting on the elements of area (each = $dF$) of the section consist of shears (whose sum = $J$, the "total shear") in the plane of the section and parallel to the force plane, and of normal stress parallel to $AK$ and proportion- al per unit of area to the distances of the $dF$'s on which they act from the neutral axis $NC'$, real or ideal (ideal in this figure). Imagine the outermost fibre $KA$, whose dis- tance from the gravity axis is $=a$ and from the neutral axis Fig. 296. =+a, to be prolonged an amount $AA'$, whose length by some arbitrary scale represents the normal stress (tension or compression) to which the $dF$ at $A$ is subjected. Then, if a plane be passed through $A'$ and the neutral axis $NC'$, the lengths, such as $mr$, parallel to $AA'$, intercepted between this plane and the section itself, represent the stress-inten- 300 FLEXURE OBLIQUE FORCES. 349 sities (i. e., per unit area) on the respective $dF$'s. (In this particular figure these stresses are all of one kind, all tensile or all compression; but if the neutral axis occurs within the section, as in the next section, there will be two kinds on the two sides of $NC'$.) Through $C'$, the point determined in $AM'NC'$ by the intercept $CC'$ of the centre of gravity, pass a plane $A'M'T'$ parallel to the section itself; it will divide the stress-intensity $AA'$ into two parts $p_1$ and $p_2$ and will enable us to express the stress-intensity $mr$, on any $dF$ at a distance $z$ from the gravity-axis $GC$, in two parts; one part the same for all $dF$'s, the other dependent on $z$, thus: \[ (\text{Stress-intensity on any } dF) = p_1 + \frac{z}{e} p_2 \quad . . . \quad (1) \] and hence the \[ (\text{actual normal stress on any } dF) = p_1 dF + \frac{z}{e} p_2 dF \quad (2) \] For example, the stress-intensity on the fibre at $T$, where \[ t = -e, \] will be \( p_1 - \frac{e_1}{e} p_2 \) and it is now seen how we may find the stress at any $dF$ when $p_1$ and $p_2$ have been found. If the distance $a$, between the neutral and gravity axes is desired, we have, by similar triangles \[ p_1 : CC' : a \] whence \( a = \frac{P_1}{p_1} e \quad . . . \quad (3) \] It is now readily seen, graphically, that the stresses or elastic forces represented by the equal intercepts between the parallel planes $AMT$ and $A'M'T'$, constitute a uniformly distributed normal stress, which will be called the "uniform thrust," or simply the thrust (or pull, as the case may be) of an intensity \( p_0 \), and \( z \) of an amount \( f p_0 dF = P_0 f dF = P_0 F \). It is also evident that the positive intercepts forming the 350 MECHANICS OF ENGINEERING. wedge $A''N''G''$ and the negative intercepts forming the $M''M'G'C'$ form a system of "graded stresses" whose combination (algebraic) with those of the "thrust" shows the two sets of forces to be equivalent to the actual system of normal stresses represented by the small prisms forming the imaginary solid $AM'T'...AM'T$. It will be shown that these graded stresses constitute a "stress-couple". Analytically, the object of this classification of the normal stresses into a thrust and a stress-couple, may be made apparent as follows: In dealing with the free body $KLM$ Fig. 296, we shall have only to sum the components, parallel to the beam, of all forces acting (external and elastic), also those $\uparrow$ to the beam; and also sum their moments about some axis $\uparrow$ to the force plane. Let this axis of moments be $GC$ the gravity-axis of the section (and not the neutral axis); also take $P_1$ and $P_2$ on the beam at $T'$ to its (any) in force-plane). Let us see what part the elastic forces will play in these three summations. See Fig. 297, which gives merely a side view. Referring to eq. (2) we see that $$\left[\text{The } \Sigma X\text{ of the forces}\right] = p_i \int_{i'}^f dF + \frac{P_2}{e} \int_{i'}^f z dF$$ i.e. " " " " " " = \(p_i F + \frac{P_2}{e} F z\) (see §4 § 23). But as the's are measured from $G$ a gravity axis, $z$ must be zero. Hence $$\left[\text{The } \Sigma X\text{ of the Elastic forces}\right] = p_i F - \left\{\begin{array}{l} \text{the thrust} \\ \text{(or pull)} \end{array}\right\}$$ (4) FLEXURE. OBLIQUE FORCES. 351 Also, (The $Z$ of the Elastic forces) = J = the shear; * (6) while for moments about G (see eq. (1)) (The $Z$ (moments) of the elastic forces) = $$\int_{i}^{p_i dF} + \int_{i}^{p_i dF} = p_i \int_{i}^{dF} + \frac{p_i}{e} \int_{i}^{dF}$$ and hence finally (The $Z$ (moments) of the Elastic Forces) = $\frac{P_i I_0}{e}$ = the moment . . . (6) where $I_0 = \int d^2F$, is the "moment of inertia" of the section about the gravity axis $G$, (not the neutral axis). The expression in (6) may be called the moment of the stress-couple, understanding by stress-couple a couple to which the graded stresses of Fig. 297 are equivalent. That these graded stresses are equivalent to a couple is shown by the fact that although they are $X$ forces they do not appear in eq. 4, for $ZZ$; hence the sum of the tensions in that set of normal stresses. We have therefore gained three advantages, that, of the three quantities $J$ (lbs.), $p_i$ (the per sq inch), and $p_i$ (lbs. per sq inch), one of which, with the form of the section, completely determines the stresses in the section, equations (4), (5), and (6) contain only one each, and hence algebraic elimination is unnecessary for finding any one of them; and that the axis of reference of the moment of inertia is the same axis of the section as was used in former problems in flexure. Another mode of stating eqs. (4), (5) and (6) is this: The sum of the components, parallel to the beam, of the external forces is balanced by the thrust or pull; those perpen- 352 MECHANICS OF ENGINEERING. diacular to the beam are balanced by the *shear*; while the sum of the moments of the external forces about the gravity axis of the section is balanced by that of the stress-couple. Notice that the thrust can have no moment about the gravity axis referred to. The Equation for Safe Loading, then, will be this: \[ (a) \quad (p_i \pm p_o) \text{ max. } = R^o \quad (7) \] (b) \quad (p_i \pm q_i) \text{ max. } = R^e \quad (8) For $R'$, see table in § 251. The double sign provides for the cases where $p_i$ and $p_o$ are of opposite kinds, one tension the other compression. Of course $(p_i + p_o)$ max is not the same thing as $(p_i \text{ max} + p_o \text{ max})$. In most cases in practice $q_i = e$, and then the part ($b$) of eq. (7) is unnecessary. 295a Elastic Curve with Oblique Forces.—By elastic curve is now meant the locus of the centres of gravity of the sections.) Since the effect of a uniform pull (or push) differs from those occurring under perpendicular forces only in the addition of a uniform thrust (or pull), whose effect on the short lengths (=dx) of fibres between two consecutive sections $U'V'$ and $U_V$, Fig. 297, is felt equally by all, the location of the centre of curvature (or) would be appreciably different from what it would be as determined by the stress-couple alone. Thus (within the elastic limit), strains being proportional to the stresses producing them, if the forces of the stress-couple acted alone, the length $dx = G_0G'$ of a small portion of a fibre at the gravity axis would remain unchanged, and the lengthening and shortening of the other fibre-lengths between the two sections $U_V$ and $U'V'$, originally parallel, would occasion the turning of $U'V'$ through a small angle (relatively to $U_V$) about $G'$, into the position which it occupies in the figure (297), and $G_0G'$ would be the radius of curvature. But the effect of the uniform pull (added to that of the couple) is to shift $U'V'$ parallel to itself into the position $UV$ and hence the radius of curvature of the FLEXURE. OBLIQUE FORCES. 353 elastic curve, of which $G_r G$ is an element, is $G_r R$ instead of $G_r R'$. But the difference between $G_r R$ and $G_r R'$ is very small, being the same, relatively, as the difference between $G_r Q$ and $G_r Q'$; for instance, with wrought-iron, even if $p_n$ the intensity of the uniform pull, were as high as 22,000 lbs. per sq. in. (see § 203) $G_r G$ would exceed $G_r Q$ by only $\frac{1}{m}$ of one per cent. ($=0.0008$); hence by using $GR$ instead of $GR$ as the radius of curvature $\rho$, an error is introduced of so small an amount as to be neglected. But from § 201, eqa. (6) and (7), $\frac{EI}{d^2y} = M$, the sum of the moments of the external forces; hence for prismatic beams under oblique forces we may still use $$\pm \frac{EI}{d^2y} \cdot (M) \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1)$$ as one form for the $\Sigma$(mom.) of the elastic forces of the section about the gravity-axis; remembering that the axis $X$ must be taken parallel to beam. 296. Oblique Cantilever with Terminal Load.—Fig. 298. Let $l =$ length. The "fixing" of the lower end of the beam is its only support. Measure $x$ along the beam from $O$. Let Fig. 298. $a$ be the gravity axis of any section and $aT_1 = x\sin a$, the length of the perpendicular let fall from $a$ on the line of action of the force $P$ (load). The flexure is so slight that $aT$ is considered to be the same as before the load is al- 354 MECHANICS OF ENGINEERING. lowed to act. [If $a$ were very small, however, it is evident that this assumption would be inadmissible, since then a large proportion of $nT$ would be due to the flexure caused by the load.] Consider $nO$ free, Fig. 299. In accordance with the preceding paragraph (see eqs. (4), (5), and (6)) the elastic forces of the section consist of a shear $J$, whose value may be obtained by writing $Y.Y=0$ whence $$J = P \sin a; \quad \ldots \quad \ldots \quad (1)$$ of a uniform thrust $=p_oP$, obtained from $2X=0$, viz: $$P \cos a - p_oF = 0 \cdot p_oF = P \cos a; \quad (2)$$ and of a stress-couple whose moment [which we may write either $\frac{dI}{dx}$ or $EI\frac{dy}{dx^2}$] is determined from $Z(mom.)=0$ or $$\frac{p_oI}{e} - Pz \sin a = 0, \text{or } \frac{p_oI}{e} = Pz \sin a \quad . . . \quad (3)$$ As to the strength of the beam, we note that the stress-intensity, $p_o$, of the thrust is the same in all sections, from $O$ to $L$ (Fig. 298), and that $p_o$ is the stress-intensity in the outer fibre, (and this is compression if $a > 0$ of Fig. 299) due to the stress-couple is proportional to $a$, hence the max. of $(p_o + p_o)$ will be in the inner outer fibre at $L$, Fig. 298, where $x$ is as great as possible, i.e., will be a compression, viz. $$[p_o + p_o]_{\max} = P\left[\frac{\cos a + l(\sin a)}{F}\right] \quad . . . \quad (4)$$ $\therefore$ the equation for Safe Loading is $$R = P\left[\frac{\cos a + l(\sin a)}{F}\right] \quad . . . \quad (5)$$ since with $e_i=e$, as will be assumed here, $(p - p_o)_{\max}$ **FLEXURE. OBLIQUE FORCES.** 355 can not exceed, numerically, $[p_+ + p_-]$ max. The stress-intensity in the outer fibre along the upper edge of the beam, being $= -p_-$ (exceeding $p_+$ will be more expressive at the upper end near $O$, since there $p_+$ is small, $x$ being small); but lower down as $x$ grows larger, $p_+$ increasing, a section may be found (before reaching the point $L$) where $p_+ = p_-$ and where consequently the stress in the outer fibre is zero, or in other words the neutral axis of that section passes through the outer fibre. In any section above that section the neutral axis is imaginary, i.e., is altogether outside the section, while below it, it is within the section, but cannot pass beyond the gravity axis. Thus in Fig. 300, $OL'$ Fig. 300. Fig. 301. is the locus of the positions of the neutral axis for successive sections, while $OL$ the axis of the beam is the locus of the gravity axes (or rather of the centres of gravity) of the sections, this latter line forming the "elastic curve" under flexure. As already stated, however, the flexure is to be but slight, and $\alpha$ must not be very small. For instance, if the deflection of $O$ from its position before flexure is of such an amount as to cause the lever-arm $OR$ of $P$ about $L$ to be greater by 10 per cent than its value ($=\sin \alpha$) before flexure, the value of $p_-$ as computed from eq. (3) (with $x=$) will be less than its true value in the same proportion. The deflection of $O$ from the tangent at $L$, by § 237, Fig. 229(a) is $d = (\text{F} \sin \alpha) P / 3EI$, approximately, putting $P \sin \alpha$ 355 356 MECHANICS OF ENGINEERING. for the $P$ of Fig. 229 ; but this very deflection gives to the other component, $P \cos a$, to the tangent at $L$, a lever arm, and consequent moment, about the gravity axis of all the sections, whence for $\Sigma (moms.) = 0$ we have, (more exactly than from eq. (5) when $x = -$) $$\frac{P}{EI} - P(\sin a)\cos a + P \cos a \cdot \frac{d}{dx} \cdot \ldots \quad (6)$$ (We have supposed $P$ replaced by its components $I$ and $J$ to the fixed tangent at $L$, see Fig. 301). But even (6) will not give an exact value for $p$, at $L$; for the lever arm of $P \cos a$, viz., $d$, is > $(P \sin a)/EI$, on account of the presence and leverage of $P \cos a$ itself. The true value of $d$ in this case may be obtained by a method similar to that indicated in the next paragraph. 297. Elastic Curve of Elongate Cantilever with Terminal Load. For variety place the cantilever as in Fig. 302, so that the deflection of any section is due to the moment of $P$ about the gravity axis of any section, $n$. We may replace $P$ by its $X$ and $Y$ components, Fig. 303, $\parallel$ and $\perp$ respectively to the fixed tangent line at $L$. The origin, $O$, is taken at the free end of the beam. Let $\theta$ angle between $P$ and $X$. For a free body on $n$, being any section, we have $\Sigma (moms.) = 0$ whence $$EI \frac{dy}{dx} = P(\cos a) - P(\sin a)x \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)$$ (See eq. (1) § 295a). In this equation the right hand member is evidently (see fig. 303) a negative quantity; this is as it should be, for $EIdy/dx^2$ is negative, the curve being concave to the axis $X$ in the first quadrant. (It must be noted that the axis $X$ is always to be taken,) to the beam, for $EIdy/dx^2$ to represent the moment of the stream-couple.) Fig. 301 Fig. 302 Fig. 303 **PLEXURE OBLIQUE FORCES.** 337 Eq. (1) is not in proper form for taking the $x$-anti-derivative of both members, since one term contains the variable $y$, an unknown function of $x$. Its integration is included in a more general case given in some works on calculus, but a special solution by Prof. Robinson, of Ohio, is here subjoined for present needs.* We thus obtain as the equation of the elastic curve in Fig. 303, $$\sqrt{\frac{P_{cos}a}{EI}}\left[\xi+c_{e}^{2}\right]\left[(\sin a)x-(\cos a)y\right]-\sin a\left[c_{e}^{2}-c_{e}^{2}y\right] \quad (3)$$ In which $c_{e}$ denotes the Napierian base $=2.71828$, an abstract number, and $q$ for brevity stands for $\sqrt{P_{cos}a+EI}$. To find the deflection $d$, we make $x=y=0$, and solve for $y$; the result is $d$. The uniform thrust at $L$ is $p_{L}=F_{cos}a \ldots \ldots (3)$ while the stress intensity $p_{e}$ in the outer fibre at $L$ is ob- *Denoting $P_{cos}a$ by $p$ and $P_{sin}a$ by $q$, eq. (1) becomes $\frac{dp}{dx}=p-p_{e} \ldots \ldots (6)$* Differentiation of (6) then gives $\frac{dp}{dx}-q\frac{dy}{dx}=p-p_{e}$. Differentiating again: whence $\frac{d^2p}{dx^2}-q\frac{d^2y}{dx^2}=p-p_{e} \ldots \ldots (7)$ Letting $\frac{dy}{dx}=u \ldots \ldots (8)$ so that $u=\frac{dp}{dx}-p_{e}$, from (6), we have $\frac{du}{dx}=\frac{dp}{dx}-p_{e}$ and $\frac{d^2u}{dx^2}=\frac{d^2p}{dx^2}-p_{e}$ $\therefore \left(\begin{array}{l}{See(7)}\\ {du=dp-p_{e}, which, multi. by dx, gives dp-du=dp-dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dx}\int du\frac{dp}{dx}=du+dp_{e}}\\ {du=const.\frac{1}{dcx}\int dcx=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{2}}} \\ {\therefore c=\sqrt{\pi C^{2}/g^{ 358 MECHANICS OF ENGINEERING. tained from the moment equation for the free body $O L$ via: $$P_x = P(\sin \alpha) - P(\cos \alpha) d$$ (4) in which $c$=distance of outer fibre from the gravity axis. The equation for safe loading is written out by placing the values of $p$, $p_0$, and $d$, as derived from equations (2), (3), and (4) in the expression $$p + p_0 = R$$ To solve the resulting equation for $P$, in case that is the unknown quantity, can only be accomplished by successive assumptions and approximations, since it occurs trans- cendentially. In case a horizontal tension-member of a bridge-truss is subjected to a longitudinal tension $P''$ (due to its position in the truss and the load upon the latter) and at the same time receives a vertical pressure $P'''$ at the middle, each half will be bent in the same manner as the cantilever in Fig. 302; $\frac{1}{2} P''$ corresponding to $P \sin \alpha$, and $P'$ to $P_0 \cos \alpha$. Constant factors we now write $y = \frac{q^2}{g^2} = \frac{q^2}{g^2} - \frac{q^2}{g^2}$ The equation required . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . To determine the constants $m$ and $n$ we set $C = q^2 g^2$, and $a = C + q^2 g^2$ we find first dp/dx, i.e. by differentiating (10) $$\frac{dp}{dx} = \frac{q^2}{g^2} \left[ \frac{dq}{dx} + q \frac{dg}{dx} \right]$$ (11) Next for $pq$ $\ldots (10)$ gives $$0 = b + m \frac{dq}{dx} + n \frac{dg}{dx}$$ i.e. $b = m \frac{dq}{dx} + n \frac{dg}{dx}$ (12) Also for $pq$ $$\frac{dp}{dx} = b + m \frac{dq}{dx} + n \frac{dg}{dx}$$ (13) with $m-n$ we have $m-n = 0$ $$p' = b + m \left[ \frac{dq}{dx} + q \frac{dg}{dx} \right]$$ (14) The equation of the curve, then, substituting (14) in (10) is $$y = p' - 2 \left[ \frac{dq}{dx} + q \frac{dg}{dx} \right]$$ (15) Substituting for $p$ and $q$ we have, $$y = p' - 2 \sqrt{\frac{q^2}{g^2}} \left[ \frac{dq}{dx} + q \frac{dg}{dx} \right]$$ (16) where, let us in (15) $$\sqrt{\frac{q^2}{g^2}} = x$$ FLEXURE. OBLIQUE FORCES. 359 298. Inclined Beam with Hinge at One End.—Fig. 304. Let $e = e_1$. Required the equation for safe loading; also the maximum shear, there being but one load, $P$, and that in the middle. The vertical wall being smooth, its reaction, A diagram showing a beam with a hinge at one end and a load \( P \) at the other end. The beam is inclined at an angle \( \alpha \). The reactions at the supports are denoted by \( H \) and \( V \). The horizontal component of the reaction at \( O \) is denoted by \( H_x \), and the vertical component by \( H_y \). The vertical force due to the load is denoted by \( V \). The horizontal force due to the load is denoted by \( V_x \). \( H_x \) at \( O \) is horizontal, while that of the hinge-pin being unknown, both in amount and direction, is best replaced by its horizontal and vertical components, \( H_x \) and \( V_x \), unknown in amount only. Supposing the flexure slight, we find these external forces in the same manner as in Prob. 1 § 37, by considering the whole beam free, and obtain \[ H = F \cot a; H_x \text{ also } = F_x \cot a; V_x = P. \] (1) For any section between \( O \) and \( B \), we have, from the free body at \( O \), Fig. 305, uniform thrust = \( p_x F = H cos a \). (2) and from \( I \) (moments) = 0, \[ \frac{pI}{e} - He sin a = 0 \] (3) and the shear = \( J = H sin a = \frac{1}{2} P cos a \). (4) The max. (\( p_1 + p_2 \)) to be found on \( OB \) is : close above \( B \), where \( x = \frac{1}{2} l_1 \) and is 360 MECHANICS OF ENGINEERING. $$= \frac{H\cos a}{F} + \frac{H\sin a}{2I} \text{ which } = P\cos a \left[ \cot a + \frac{le}{2k^2} \right] \quad (5)$$ In examining sections on $CB$ let the free body be $Cw'$, Fig. 306. Then from $\Sigma$ (longitud. comp.) = 0 (the thrust =) $p_i F = V_s \sin a + H_s \cos a$ (6') i.e. $$p_i F = P[\sin a + \frac{\gamma}{2} \cos a \cot a] \quad (6)$$ while, from $\Sigma$ (moms.,$_u$) = 0, $$\frac{P_i F}{e} = V_s x' \cos a - B_s x' \sin a \quad (7')$$ i.e. $$\frac{P_i F}{e} = \frac{\gamma}{2} P \cos a x' \quad (7)$$ Hence $(p_i + p_b)$ for sections on $CB$ is greatest when $x'$ is greatest, which is when $x' = \frac{\gamma}{2} l$, $x'$ being limited between $x' = 0$ and $x' = \frac{\gamma}{2} l$, and is $(p_i + p_b)$ max. on $CB = P\cos a \left[ \tan a + \frac{\gamma}{2} \cot a + \frac{\gamma}{2} l e \right]$ (8) which is evidently greater than the max. $(p_i + p_b)$ on $BO$; see eq. (5). Hence the equation for safe loading is $$R' = P\cos a \left[ \tan a + \frac{\gamma}{2} \cot a + \frac{\gamma}{2} l e \right] \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)$$ in which $R'$ is the safe normal stress, per square unit, for the material. The shear, $J$, anywhere on $CB$, from $\Sigma$ (transverse comp.) = 0 in Fig. 306, is $$J = V_s \cos a - H_s \sin a = \frac{\gamma}{2} P\cos a \quad : (10)$$ 361 **FLEXURE. OBLIQUE FORCES.** As showing graphically all the results found, moment, thrust, and shear diagrams are drawn in Fig. 307, and also a diagram whose ordinates represent the variation of $(p_1 + p_2)$ along the beam. Each ordinate is placed vertically under the gravity axis of the section to which it refers. 298. Numerical Example of the foregoing—Fig. 308. Let the beam be of wrought iron, the load $P = 1,800$ lbs., hanging from the middle. Cross-section rectangular 2 in. by 1 in., the 2 in. being parallel to the force-plane. Required the max. normal stress in any outer fibre ; also the max. total shear. This max. stress-intensity will be in the outer fibres in the section just below $B$ and on the upper side, according to § 296, and is given by eq. (8) of that article; in which, see Fig. 308, we must substitute (inch-pound-second-system) $P = 1,800$ lbs.; $F = 2$ sq. in.; $l = \sqrt{20^2 + 1^2} = 120.6$ in.; $e = 1$ in., $T = \frac{1}{\pi} \cdot \frac{b^2}{h} = \frac{1}{\pi}$ biquad. inches; $\cot a = \frac{1}{\tan a} = 0.0996$; and $\tan a = 10$. $\therefore$ max. $(p_1 + p_2) = 1800 \times 0.0996 \left[ \frac{120.6 + e}{\pi} + \frac{e}{\pi} \cdot \frac{b^2}{h} \right]$ = 9000 lbs. per sq. inch, very nearly, compression. This is in Fig. 308. Fig. 309. Fig. 310. 362 MECHANICS OF ENGINEERING. the upper outer fibre close under $B$. In the lower outer fibre just under $B$ we have a tension = $p_1 - p_2 = 7,500$ lbs. per sq. in. (It is here supposed that the beam is secure against yielding sideways.) 300. Strength of Hooks.—An ordinary hook, see Fig. 309, may be treated as follows: The load being = $P$, if we make a horizontal section at $AB$, whose gravity axis $g$ is the one, of all sections, furthest removed from the line of action of $P$, and consider the portion $C$ free, we have the shear = $J = zero$ \begin{align*} & \text{the uniform pull} = p_1 F = P \\ & \text{while the moment of the stress-couple, from } Z (\text{moments}) = 0, \end{align*} is $$\frac{p_1 I}{c} = Pa$$ (8) For safe loading $p_1 + p_2$ must = $R'$, i.e. $$R' = P \left[ \frac{1}{F^+} \frac{ae}{I} \right]$$ (4) It is here assumed that $e = e_a$ and that the maximum $(p_1 + p_2)$ occurs at $AB$. 301 Crane.—As an exercise let the student investigate the strength of a crane, such as is shown in Fig. 310. FLEXURE LONG COLUMNS 363 CHAPTER VI FLEXURE OF "LONG COLUMNS." 302. Definitions.—By "long column" is meant a straight beam, usually prismatic, which is acted on by two compressive forces, one at each extremity, and whose length is so great compared with its diameter that it gives way (or "flows") under the action of these forces without fracture, or by crushing or splitting like a short block (see § 200). The pillars or columns used in buildings, the compression members of bridge-trusses and roofs, the "bents" of a trestle work, and the piston-roads and connecting-rods of steam-engines, are the principal practical examples of long columns. That they should be weaker than short blocks of the same material and cross-section is quite evident, but their theoretical treatment is much less satisfactory than in other cases of flexure, experiment being very largely relied on in determining the form of the formulae which theory introduces in the formulae referring to them, but even to modify the algebraic form of those formulae, thus rendering them to a certain extent empirical. 303. End Conditions.—The strength of a column is largely dependent on whether the ends are free to turn, or are fixed and thus incapable of turning. The former condition is attained by rounding off the ends, or by fitting them with hinges or ball-and-socket-joints; the latter by facing off each end to an accurate plane surface, the bearing on which it rests being plane also, and incapable of turning. A page from a book about structural engineering. 364 MECHANICS OF ENGINEERING. In the former condition the column is spoken of as having round ends; Fig. 311, (a); in the latter as having fixed ends, (or flat bases ; or square ends), Fig. 311, (b). A diagram showing a column with rounded ends and another with fixed ends. A diagram showing a column with rounded ends and another with fixed ends. A diagram showing a column with rounded ends and another with fixed ends. Sometimes a column is fixed at one end while the other end is not only round but incapable of lateral deviation from the tangent line of the other extremity; this state of end conditions is often spoken of as "Pin and Square," Fig. 311, (c). If the rounding of the ends is produced by a hinge or "pin joint," Fig. 312, both pins lying in the same plane and having immovable bearings at their extremities, the column is to be considered as round-ended as regards flexure in its own plane, but as having fixed ends as regards flexure in the plane containing the axes of the pins. The "moment of inertia" of the section of a column will be understood to be referred to a gravity axis of the section which is $\gamma$ to the plane of flexure (and this corresponds to the "force-plane" spoken of in previous chapters), or plane of the axis of column when bent. 303a. Euler's Formula.—Taking the case of a round-ended column, Fig. 313 (a), assume the middle of the length as an origin, with the axis $X$ tangent to the elastic curve at that point. The flexure being slight, we may use the form $EI \frac{d^2y}{dx^2}$ for the moment of the stress-couple in any A diagram showing a column with rounded ends and another with fixed ends. FLEXURE. LONG COLUMNS. 363 Fig. 81A. Fig. 81B. section $a$, remembering that with this notation the axis $X$ must be $l$ to the beam, as in the figure (313). Considering the free body $nC$, Fig. 313 (b), we note that the shear is zero, that the uniform thrust = $P$, and that $\mathcal{L}(\text{moments})=0$ gives ($a$ being the deflection at $O$) $$\frac{d\mathcal{L}}{dx}=P(a-y) \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1)$$ Multiplying each side by $dx$ we have $$EI \frac{d^2y}{dx^2}dy=P(a-y)dy \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (2)$$ Since this equation is true for the $y$, $dx$, $dy$, and $d\mathcal{L}$ of any element of arc of the elastic curve, we may suppose it written out for each element from $O$ where $y=0$, and $dy=0$, up to any element, (where $dy=dy$ and $y=y$) (see Fig. 314) and then write the sum of the left hand members equal to that of the right hand members, remembering that, since $dx$ is assumed constant, $1+dx$ is a common factor on the left. In other words, integrate between $O$ and any point of the curve, $a$. That is, $$EI\int_{O}^{a}(dy)[d(y)]dy=Pa\int_{O}^{a}dy-P\int_{O}^{a}ydy \quad \quad (3)$$ The product $dy dy$ has been written $(dy)(dy)$, (for $dy$ is 366 MECHANICS OF ENGINEERING. the differential or increment of $dy$) and is of a form like $dx$, or $dy$. Performing the integration we have $$\frac{EI}{d^2} \cdot \frac{dy}{2} = P \cdot \frac{y^2}{2} \cdot \ldots \quad (4)$$ which is in a form applicable to any point of the curve, and contains the variables $x$ and $y$ and their increments $dx$ and $dy$. In order to separate the variables, solve for $dz$, and we have $$dz = \sqrt{\frac{EI}{P}} \cdot \frac{dy}{2\sqrt{2y-y^2}} \text{ or } dz = \sqrt{\frac{EI}{P}} \cdot \frac{d(y)}{\sqrt{2\frac{a}{a}-\left(\frac{y}{a}\right)^2}} \quad (5)$$ $$\therefore dz = \pm \sqrt{\frac{EI}{P}} \int_0^z \frac{d(y)}{\sqrt{2\frac{a}{a}-\left(\frac{y}{a}\right)^2}}$$ i.e., $z = \pm \sqrt{\frac{EI}{P}} (\text{vers. sin} -\frac{y}{a}) \quad (6)$$ (6) is the equation of the elastic curve $DOC$, Fig. 313(a), and contains the definition $a$. If $P$ and $a$ are both given, $y$ can be computed for a given $x$, and vice versa, and thus the curve traced out, but we would naturally suppose it to depend on $P$, for in eq. (6) when $x = \frac{x}{4}$, $y$ should equal $\frac{x}{4}$. Making these substitutions we obtain $$y_1 = \pm \sqrt{\frac{EI}{P}} (\text{vers. sin} -1.00); i.e., y_1 = \pm \sqrt{\frac{EI}{P}} \cdot \frac{\pi}{2}$$ (7) Since $a$ has vanished from eq. (7) the value for $P$ obtained from this equation, viz. $$P_0 = EI \cdot \frac{n^2}{F} \quad (8)$$ is independent of $a$, and is : to be regarded as that force (at each end of the rounded column in Fig. 313) which will hold the column at any small deflection at which it may previously have been set. FLEXURE LONG COLUMNS. 367 In other words, if the force is less than $P_0$ no flexure at all will be produced, and hence $P_0$ is sometimes called the force producing "incipient flexure." [This is roughly verified by exerting a downward pressure with the hand on the upper end of the flexible rod (a T-square-blade for instance) placed vertically on the floor of a room ; the pressure must reach a definite value before a decided buckling takes place, and then a very slight increase of pressure occasions a large increase of deflection.] It is also evident that a force slightly greater than $P_0$ would very largely increase the deflection, thus gaining for itself a great lateral arm about the middle section as to cause rupture. For this reason eq. (8) may be looked upon as giving the Breaking Load of a column with round ends, and is called Euler's formula. Referring now to Fig. 311, it will be seen that if the three parts into which the flat-ended column is divided by its two points of inflection $A$ and $B$ are considered free, individually, in Fig. 315, the forces acting will be as there shown, viz: At the points of inflection there is no stress-couple, i.e., $\sigma_{AB} = \sigma_{BC} = 0$, and hence $P_0 = P$, and hence the portion $AB$ is in the condition of a round-ended column. Also, the tangents to the elastic curves at $O$ and $C$ being preserved vertical by the frictionless guide-blocks and guides (which are introduced here simply as a theoretical method of preventing the ends from turning but do not interfere with vertical freedom) $OA$ is in the same state of flexure as half of $AB$ and hence has equal force. Hence the length $AB$ must = one half the total length $l$ of the flat-ended column. In other words, the breaking load of a round-ended column of length = $\frac{1}{2} l$, is in the same as that of a flat-ended column of length = $l$. Hence for the $l$ of eq. (8) write $\frac{1}{2} l$, and we have as the breaking load of a column with flat-ends and of length = $l$. A diagram showing a flexible rod with points A and B where it bends over. Fig. 315 368 MECHANICS OF ENGINEERING. $$P_1 = 4 \frac{EI}{\pi^2} \quad \ldots \quad (9)$$ Similar reasoning, applied to the "pin-and-square" mode of support (in Fig. 311) where the points of inflection are at $R$, approximately $\frac{l}{2}$ from $C$, and at the extremity $O$ itself, calls for the substitution of $\frac{l}{2}$ for $l$ in eq. (8), and hence the breaking load of a "pin-and-square" column, of length $l$, is $$P_s = \frac{9}{4} \frac{EI}{\pi^2} \quad \ldots \quad (10)$$ Comparing eqs. (8), (9), and (10), and calling the value of $P_1$ (flat-ends) unity, we derive the following statement: The breaking loads of a given column are as the numbers
1 9/16 $\frac{\pi}{4}$ according to the flat-ends
flat-ends pin-and-square round-ends
mode of support. These ratios are approximately verified in practice. Euler's Formula (i.e., eq. (8) and those derived from it, (9) and (10)) when considered as giving the breaking load is peculiar in this respect, that it contains no reference to the stress per unit of area necessary to rupture the material of the column, but merely assumes that the load producing "incipient flexure", i.e., which produces any bending at all, will eventually break the bar because of its greater and greater tendency to bend for a given load. In the cantilever of Fig. 341 the bending of the beam does not sensibly affect the lever-arm of the load about the wall-section, but with a column, the lever-arm of the load about the mid-section is almost entirely due to the deflection produced. 394. Example. Euler's formula is only approximately verified by experiment. As an example of its use when considered as giving the force producing "incipient flexure" it will now be applied in the case of a steel T-square-blade whose ends are free to turn. Hence we use the round-end formula eq. (8) of §303, with the modulus of elasticity $E = 30,000,000$ lbs. per sq. inch. The dimensions A diagram showing a T-square-blade with one end fixed and the other end free to rotate. FLEXURE. LONG COLUMNS. 369 are as follows: the length $l = 30$ in., thickness $= \frac{1}{8}$ of an inch, and width $= 2$ inches. The moment of inertia, $I$, about a gravity axis of the section $\parallel$ to the width (the plane of bending being $\parallel$ to the thickness) is $(\S 247)$ $$I = \frac{1}{12} W^2 = \frac{1}{12} \times \left(\frac{1}{30}\right)^2 = \frac{1}{360} = 162,000 \text{ biquad. inches}.$$ $\therefore$, with $\pi = 22 + 7,$ $$P_0 = E I = \frac{\pi^2}{F} = \frac{30,000,000}{162,000} \cdot \frac{22}{7} \cdot \frac{1}{900} = 2.03 \text{ lbs.}$$ Experiment showed that the force, a very small addition to which caused a large increase of deflection or side-buckling, was about 2 lbs. 305. Hodgkinson's Formula for Columns.—The principal practical use of Euler's formula was to furnish a general form of expression for breaking load, to Eaton Hodgkinson, who experimented in England in 1840 upon columns of iron and timber. According to Euler's formula we have for cylindrical columns, $I$ being $\frac{\pi d^4}{64}$, $E$ is $\frac{d^4}{F}$ (§ 247), for flat-ends . . . . . $P_1 = \frac{1}{16} E I \cdot \frac{d^4}{F}$ i.e., proportional to the fourth power of the diameter, and inversely as the square of the length. But Hodgkinson's experiments gave for wrought-iron cylinders $$P_1 = (\text{const}) \times \frac{d^4}{F},$$ and for cast iron $P_1 = (\text{const}) \times \frac{d^4}{F - l}$. Again, for a square column, whose side $= b$, Euler's formula would give $$P_1 = \frac{b^4}{8} \times \frac{d^4}{F}$$ while Hodgkinson found for square pillars of wood $$P_1 = (\text{const}) \times \frac{b^4}{F}$$ 370 MECHANICS OF ENGINEERING. Hence in the case of wood these experiments indicated the same powers for $b$ and $l$ as Euler's formula, but with a different constant factor; while for cast and wrought iron the power was found to be $\frac{1}{3}$. Hodgkinson's formulae are as follows, and evidently not homogeneous; the prescribed units should . . . be carefully followed. $d$ denotes the diameter of the cylindrical columns, $b$ the side of square columns, $l$—length.
For solid cylindrical cast iron columns, flat-ends; $\frac{1}{3} \times (d \text{ in inches})^{\frac{1}{2}} \times (l \text{ in ft.})^{\frac{1}{2}}$
Breaking load in tons of 2,240 lbs. each
For solid cylindrical wrought iron columns, flat-ends; $\frac{1}{3} \times (d \text{ in inches})^{\frac{1}{2}} \times (l \text{ in ft.})^{\frac{1}{2}}$
Breaking load in tons of 2,240 lbs. each
For solid square columns of dry oak, flat-ends; $\frac{1}{3} \times (b \text{ in inches})^{\frac{1}{2}} \times (l \text{ in ft.})^{\frac{1}{2}}$
Breaking load in tons of 2,240 lbs. each
For solid square columns of dry fir, flat-ends; $\frac{1}{3} \times (b \text{ in inches})^{\frac{1}{2}} \times (l \text{ in ft.})^{\frac{1}{2}}$
Breaking load in tons of 2,240 lbs. each
Hodgkinson found that when the mode of support was "pin-and-square," the breaking load was about $\frac{1}{2}$ as great; and when the ends were rounded, about $\frac{1}{3}$ as great as with flat ends. These ratios differ somewhat from the theoretical ones mentioned in §803, just after eq. (10). Experiment shows that, strictly speaking, pin-ends are not equal to pin-and-square; enmity causes additional strength; for the friction of the pins in their bearings hinders the turning of the ends somewhat. As the lengths become smaller the value of the breaking load in Hodgkinson's formula increases rapidly, until it becomes larger than would be obtained by using the formula for the crushing resistance of a short block ($801$ viz., $FC$, i.e., the sectional area $\times$ the crushing resistance per unit of area). In such a case the pillar is called a short column, or "short block," and the value $FC$ is to be taken as the breaking FLEXURE. LONG COLUMNS. 371 load. This distinction is necessary in using Hodgkinson's formulae; i.e., the breaking load is the smaller of the two values, $FC$ and that obtained by Hodgkinson's rule. In present practice Hodgkinson's formulae are not often used except for hollow cylindrical iron columns, for which with $d$, and $d_0$ the external and internal diameters, we have for flat-ends $$\text{Breaking load in tons} = \text{Const.} \times (d_0 \text{in.}^{\frac{3}{2}} - d \text{in.}^{\frac{3}{2}})$$ of 2,240 lbs. each $\left\{\right.$ = Const. $\times (d_0 \text{in.}^{\frac{3}{2}} - d \text{in.}^{\frac{3}{2}})$ $\left.\begin{array}{l}{(f \text{ in feet})}\\ {= 44.16 \text{ for cast iron, and } 134 \text{ for wrought, while } n = 1.7 \text{ for cast-iron and } n = 2 \text{ for wrought.}}\end{array}\right\}$ 306. Examples of Hodgkinson's Formula.--Example 1. Required the breaking weight of a wrought-iron pipe used as a long column, having a length of 12 feet, an internal diameter of 3 in., and an external diameter of 3$\frac{1}{4}$ inches, the ends having well fitted flat bases. If we had regard simply to the sectional area of metal, which is $A = \pi d_0^2/4$, we would estimate the column as a short block (or short column) we should have for its compressive load at the elastic limit (see table §303) $P'' = FC''$ $$= 1.92 \times 24,000 = 39,890 \text{ lbs. and the safe load } P' \text{ may be taken at } 16,000 \text{ lbs.}$$ But by the last formula of the preceding article we have $$\text{Breaking load in tons} = 134.0 \times \frac{(3.25)^{\frac{3}{2}} - (3)^{\frac{3}{2}}}{12} = 15.07 \text{ tons}$$ i.e.-15.07 × 2240 = 33,768 lbs. Detail $\left\{\log(3.25)\right\} \times 355 = 0.511883 \times 355 = 1.817184$ $$[\log(3.00)\right] \times 355 = 0.477,121 \times 355 = 1.693,779;$$ and the corresponding numbers are 65.6 and 49.4; their difference = 16.2; hence $$\text{Br. load in long tons} = \frac{134 + 16.9}{144} = 15.072 \text{ long tons}.$$ 372 MECHANICS OF ENGINEERING. With a "factor of safety" (see § 205) of four, we have, as the safe load, $P = 8,442$ lbs. This being less than the 16000 lbs. obtained from the "short block" formula, should be adopted. If the ends were rounded the safe load would be one-third of this i.e., would be 2,814 lbs.; while with pin-and-square end-conditions, we should have one-half, or 4,211 lbs. Example: We wish to determine the safe load for a given a solid cylindrical cast-iron pillar with flat ends that its safe load may be 13,440 lbs. taking $d$ as a factor of safety. Let $d =$ the unknown diameter. Using the proper formula in § 305, and hence expressing the breaking load, which is to be six times the given safe load, in long tons we have (the length of column being 16 ft.) $$\frac{13440 \times 6}{16} = 44.16 \text{ (in inches)}^{16} \cdot \cdot \cdot (1)$$ i.e. [d in inches]^{16} = 36 × 16^{16} (2) or $\log d = \frac{1}{16} (\log 36 + 1.7 \times \log 16 - \log 44.16)$ $\therefore \log d = \frac{1}{16}(1.958278) - 0.551627 \therefore d = 3.56$ ins. This result is for flat ends. If the ends were rounded, we should obtain $d = 4.85$ inches. 307. Rankine's Formula for Columns.—The formula of this name (sometimes called Gordon's, in some of its forms) has a somewhat more rational basis than Euler's, in that it introduces the maximum normal stress in the outer fibre and is applicable to a column or block of any length, but still contains assumptions not strictly borne out in theory, thus introducing some considerable error in experimental determination. It may be developed as follows: Since the flat-ended columns in Fig. 315 the middle portion $AB$, between the inflection points $A$ and $B$, is acted on at each end by a thrust $P$, not accompanied by any shear or stress-couple, it will be simpler to treat that A diagram showing a column with flat ends and a thrust applied at each end. 372 FLEXURE LONG COLUMNS. 373 portion alone Fig. 316, (a), since the thrust and stress-couple induced in the section at $R$, the middle of $AB$, will be equal to those at the flat ends, $O$ and $C$, in Fig. 315. Let $\alpha$ denote the deflection of $B$ from the straight line $AB$. Now consider the portion $AB$ as a free body in Fig. 316, (b), putting in the stress-couple, which may be classified into a uniform thrust = $p_1 F$, and a stress couple of moment = $\frac{p_1 I}{c}$, (see § 294). (The shear is evidently zero, from $\Sigma$ (hor comp.) = 0). Here $p_1$ denotes the uniform pressure (per unit of area), due to the uniform thrust, and $p_2$ the pressure or tension (per unit of area), in the elastic forces constituting the stress-couple, on the element $AB$. $\gamma$ is the distance between the gravity axis ("to plans of flexure") of the section. $F$ is the total area of the section. $I$ is the moment of inertia about the said gravity axis, $g$ $$\Sigma \text{(vert. comp.)} - 0 \text{ gives } P = p_1 F \quad . . . \quad (1)$$ $$\Sigma \text{(moments)} - 0 \text{ gives } Pu = \frac{p_1 I}{c} \quad . . . \quad (2)$$ For any section, $s$, between $A$ and $R$, we would evidently have the same $p_1$ as at $R$, but a smaller $p_2$, since $Py < Pa$ while $e$, $I$, and $F$, do not change, the column being prismatic. Hence the max. $(p_1 + p_2)$ is on the concave edge at $R$ and for safety should be no more than $C + s$, where $C$ is the Modulus of Crushing (§ 201) and $s$ is a "factor of safety." Solving (1) and (2) for $p_1$ and $p_2$, and putting their sum = $C + s$, we have $$\frac{P + Pac}{F^+} = \frac{C}{I} \quad . . . \quad . . . \quad (8)$$ We might now solve for $P$ and call it the safe load, but it A diagram showing a flexure long column with sections A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R. (a) (b) (c) (d) (e) (f) (g) (h) (i) (j) (k) (l) (m) (n) (o) (p) (q) (r) Fig. 86. 373 374 MECHANICS OF ENGINEERING. is customary to present the formula in a form for giving the breaking load, the factor of safety being applied afterward. Hence we shall make $e = 1$, and solve for $P$, calling it then the breaking load. Now the deflection is unknown, but may be expressed approximately, as follows, in terms of $e$ and $l$. Suppose two columns of lengths $= l'$ and $l''$, each bearing its safe load. Then at the point $\frac{EF}{p'} + \frac{EF}{e'}$; i.e., $$E' = p' p''.$$ Considering the curve $AB$ as a circular arc we have (see § 200) $a' - l'' = 52 e'$, i.e. $\frac{52 E'}{p'} = e'$, and similarly for the other column, $a'' = \frac{52 E''}{p''}$. If the columns are of the same material $E' = E''$, and if each is bearing its safe load we may assume $p' = p''$ nearly, in which case the term $p' + E'' = p' + E'$, and we may say that the deflection $a$, under safe load, is proportional to $(length)^2 + e$, approximately, i.e., that $ae = \beta l$, where $\beta$ is a constant (see § 86), depending on experiment and different for different materials, and $l$ the full length. We may also write, for convenience, $f = \frac{1}{k}$, $k$ being the radius of gyration (see § 85). Hence, finally, we have from eq. (3) Breaking load $\{\} = P_i = \frac{FC}{1 + \beta^2 k^2} \cdot \cdot \cdot \cdot \cdot \cdot (4)$ This is known as Rankine's formula. By the same reasoning as in § 303, for a round-ended column we substitute $2/l$ for $l$; for a pin-and-square column $\frac{1}{4}$ for $l$; and obtain Breaking load $\{\} = P_i = \frac{FC}{1 + 4\beta^2 k^2} \cdot \cdot \cdot \cdot \cdot \cdot (5)$ Breaking load for a pin-and-square column $\{\} = P_i = \frac{FC}{1 + \frac{1}{4}\beta^2 k^2} \cdot \cdot \cdot \cdot \cdot (6)$ **FLEXURE LONG COLUMNS** 375 These formulae, (4), (5), and (6), unlike Hodgkinson's, are of a homogeneous form. Any convenient system of units may therefore be used in them. Rankine gives the following values for $C$ and $\beta$, to be used in these formulae. These are based on Hodgkinson's experiments.
CAST IRON W'n't IRON TIMBER
$C$ in lbs. per sq. in. 80,000 36,000 7,200
$\beta$ (abstract number) 1
6,400
1
36,000
1
3,000
If these numerical values of $C$ are used $F$ must be expressed in sq. inches and $P$ in Pounds. Rankine recommends 4 as a factor of safety for iron in quiescent structures, 5 under moving loads; 10 for timber. The N. J. Iron & Steel Co. use Rankine's formula for their wrought iron rolled beams, when used as columns, with a factor of safety of 4½%. 308. Examples Using Rankine's Formula.--EXAMPLE 1.-- Take the same data for a wrought iron pipe used as a column, as in example I, § 306; i.e., $l = 12$ ft = 144 inches, $F = \frac{1}{4}(3\frac{3}{4} - 2\frac{3}{4}) = 1.227$ sq. inches, while $k'$ for a narrow circular ring like the present section may be put $= \frac{1}{2}\cdot1\frac{1}{2}$ (see § 98) sq. inches. With these values, and $C = 36,000$ lbs. per sq. in., and $\beta = \frac{1}{\sqrt{3}}$ (for wrought iron), we have from eq. (4), for flat ends, $$P_1 = \frac{1.227 \times 36,000}{1 + \frac{1}{36,000} \cdot \frac{144^2}{(1.625)^2}} = 30743.6 \text{ lbs.}$$ (1) This being the breaking load, the safe load may be taken $= \frac{1}{4}$ or $\frac{1}{8}$ of 30743.6 lbs., according as the structure of 376 MECHANICS OF ENGINEERING. which the column is a member is quiescent or subject to vibration from moving loads. By Hodgkinson's formula 33,708 lbs. was obtained as a breaking load in this case (§ 306). For rounded ends we should obtain (eq. 5) $$P_c = 16,100 \text{ lbs., as break. load} \quad \text{(3)}$$ and for pin-and-square, eq. (6) $$P_p = 24,908 \text{ lbs., as break. load} \quad \text{(3)}$$ **Example 2** - (Same as Example 1, § 306). Required by Rankine's formula the necessary diameter, $d$, to be given a solid cylindrical cast-iron pillar, 16 ft. in length, with rounded ends, that its safe load may be six tons (i.e., of 2,240 lbs. each) taking 6 as a factor of safety. $$P = \frac{\pi d^2}{4} \quad \text{(4)}$$ while the value of $k$ is thus obtained. From § 247, $I$ for a full circle about its diameter $\frac{\pi d^2}{4}$ is $k^2 I = k^2 \frac{\pi d^2}{4}$. Hence eq. (5) of § 307 becomes $$P_o = \frac{\frac{\pi d^2}{4} Q}{1 + 4\alpha \frac{16r}{d}} \quad \text{(1)}$$ where $Q$ is the breaking load is to be $= 6 \times 2,240$ lbs., $O$ for cast-iron is $80,000$ lbs. per sq. inch, while $k$ (abstract number) = $\frac{c}{d}$ Solving for $d$ we have the biquadratic equation: $$d^2 - \frac{28 \times 6 \times 6 \times 2,240}{22 \times 80,000} d - \frac{28 \times 6 \times 6 \times 2,240 \times 16^2 \times 4}{22 \times 80,000 \times 400} = 0$$ whence $d = 0.641 (1 + 33.99)$, and taking the upper sign, finally, $d = \sqrt{22.4} = 4.73$ inches. (By Hodgkinson's rule we obtained 4.85 inches). 309. Radius of Gyration. - The following table, taken from p. 523 of Rankine's Civil Engineering, gives values of $P$, the square of the least radius of gyration of the given cross-section about a gravity-axis. By giving the least value of
P$d^2$
16,100 lbs.16
24,908 lbs.19
33,708 lbs.21
we find that for a square section with sides equal to the radius of gyration, the breaking load is approximately twice that for a circular section of the same radius. A diagram showing a square section with sides equal to the radius of gyration. The following table shows the breaking loads for various sections when the radius of gyration is constant.
Section$d^2$P
Square1616,100 lbs.
Circular1924,908 lbs.
Round2133,708 lbs.
It will be seen that the breaking load increases rapidly with increase in radius of gyration. FLEXURE LONG COLUMNS 377 It is implied that the plane of flexure is not determined by the end-conditions of the column; (i.e., it is implied that the column has either flat ends or round ends.) If either end (or both) is a pin-joint the column may need to be treated as having a flat-end as regards flexure in a plane containing the axis of the column and the axis of the pin, if the bearings of the pin are firm; while as regards flexure in a plane perpendicular to the pin it is to be considered round-ended at that point. In the case of "thin cell" the value of $k$ is strictly true for metal infinitely thin and of uniform thickness; still, if that thickness does not exceed $\frac{1}{6}$ of the exterior diameter, the form given is sufficiently near for practical purposes; similar statements apply to the branching forms. (a) A rectangular section with dimensions labeled. (b) A square section with dimensions labeled. (c) A circular section with dimensions labeled. (d) A circular section with a pin-joint. (e) A rectangular section with a pin-joint. Fig. 317. (f) A rectangular section with a pin-joint. (g) A square section with a pin-joint. (h) A circular section with a pin-joint. (i) A circular section with a pin-joint. (j) A rectangular section with a pin-joint. Fig. 318. Solid Rectangle. $h =$ least side. Thin Square Cell. Side = $h$. Thin Rectangular Cell. $h =$ least side. Solid Circular Section. Diameter = $d$. Thin Circular Cell. Exterior diam. = $d$. Angle-Iron of Equal ribs Fig. 317(a). $k = \frac{1}{h} h$ Fig. 317(b). $k = \frac{1}{h} h^2$ Fig. 317(c). $k = \frac{1}{h} h + \frac{3b}{12} h + b$ Fig. 317(d). $k = \frac{1}{h} d$ Fig. 317(e). $k = \frac{1}{h} d^2$ Fig. 317(f). $k = \frac{1}{h} d^2$ 378 MECHANICS OF ENGINEERING. Angle-Iron of unequal ribs. Fig. 318 (a). $k = \frac{F}{A} = \frac{P}{12(1 + M)}$ Cross of equal arms. Fig. 318 (b). $k = \frac{P}{A} = \frac{P}{12}$ I-Beam as a pillar. Let area of web = B. " " both flanges = A. Channel. Fig. 318 (d). $k = \frac{A}{B} = \frac{A}{12(A + B)} + \frac{AB}{4(A + B)}$ Iron. Let area of web = B, of flanges = A (both). $h$ extends from edge of flange to middle of web. Illustration of a channel iron with flanges. Illustration of a channel iron with flanges. 310. Built Columns.—The "compression members" of wrought-iron bridge trusses are generally composed of several pieces riveted together, the most common forms being the Phoenix column, ring-shape, in segments, and combinations of channels, plates, and lattices, some of which are shown in Figs. 319 and 320. Experiments on full size columns of these kinds were made by the U. S. Testing Board at the Watertown Arsenal about 1880. The Phoenix columns ranged from 8 in. to 28 feet in length, and from 1 to 42 in the value of the ratio of length to diameter. The breaking loads were found to be somewhat in excess of the values computed by Baehr's formula, ranging from 10 to 40 tons per square foot. In this pocket-book issued by the Phoenix company they give the following formula for their columns, (wrought-iron.) Illustration of a Phoenix column. Illustration of a Phoenix column. FLEXURE. LONG COLUMNS. 379 Breaking load in lbs. $I = \frac{50,000 F}{1 + \frac{F}{h}}$ . . . . . . . . (1) for flat-ended columns $= \frac{50,000 F}{1 + \frac{F}{h}}$ . . . . . . . . (1) where $F =$ area in sq. in., $l =$ length, and $h =$ external diameter, both in the same unit. Many different formulae have been proposed by different engineers to satisfy these and other recent experiments on columns, but all are of the general form of Rankine's. For instance Mr. Bouscaren, of the Keystone Bridge Co., claims that the strength of Phoenix columns is best given by the formula Breaking load in lbs. for flat-ends $= \frac{38,000 F}{1 + 100,000 F}$ . . . . . . . . (2) (F must be in square inches.) The moments of inertia, $I$, and thence the value of $k' = I + F$, for such sections as those given in Figs. 319 and 320 may be found by the rules of §§ 85-95, (see also § 258). 311. Moment of Inertia of Built Column. Example.-It is proposed to form a column by joining two I-beams by lattice-work, Fig. 321(a). (While the lattice-work is relied upon to cause the beams to act together as one piece, it is not regarded in estimating the area $F$, or the moment of inertia, of the cross section.) It is also required to find the proper distance apart $x$, Fig. 321, at which these beams must be placed, from centre to centre, so that the liability to flexure shall be equal for all axial planes, i.e., that the sum of the moments of inertia shall be the same about all gravity axes. This condition will be fulfilled if $f_r$ can be made $= I_x * (\text{§} 89)$. $O$ being the centre of gravity of the compound section, and $X$ perpendicular to the parallel webs of the two equal I-beams. Let $F^o =$ sectional area of one of the 1-beams, $I_x$ (see Fig. 321(a)) its moment of inertia about its web-axis $I_x'$ that about an axis $\parallel$ to web. (These quantities can be) *That is, with flat ends or ball ends ; but with pin ends, Fig. 812, if the pin is $\parallel$ to $X$, put $4I_x = I_{xx}$; if $\parallel$ to $Y$, put $4I_y = I_{yy}$. 380 MECHANICS OF ENGINEERING. found in the hand-book of the iron company, for each size of rolled beam). Then the total $I_x = 2F_x$; and total $I_r = 2\left[F_r + F'\left(\frac{x}{3}\right)^2\right]$ (see §8 eq. 4.) If these are to be equal, we write them so and solve for $x$, obtaining $x = \sqrt{\frac{4(F'_r - F_r)}{F'}}$ . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) 312. Numerically: suppose each girder to be a 10½ inch light I-beam, 105 lbs. per yard, of the N. J. Steel and Iron Co., in whose hand-book we find that for this beam $I'_r = 185.6$ biquad. inches, and $I_r = 9.43$ biquad. inches, while $F' = 10.44$ sq. inches. With these values in eq. (1) we have $x = \sqrt{\frac{4(185.6 - 9.43)}{10.44}} = \sqrt{67.5} = 8.21$ inches. (a) and (b) diagrams of I-beams. The square of the radius of gyration will be $k^2 = 2F_x + 2F' = 371.2 + 20.88 - 17.7$ sq. in. . . . . . . . . . . . . . . (2) and in the same for any gravity axis (see §8). As an additional example, suppose the two I-beams united by plates instead of lattice. Let the thickness of the plate $= t$, Fig. 321, (b). Neglect the rivet-holes. The distance $a$ is known from the hand-book. The student may derive a formula for $x$, imposing the condition that (total $I_x$) = $I_r$ Fig. 321 FLEXURE. LONG COLUMNS. 381 §13. Trussed Girders.—When a horizontal beam is trussed Diagram showing a trussed girder with a vertical plane of flexure. Fig. 322 in the manner indicated in Fig. 322, with a single post or strut under the middle and two tie-rods, it is subjected to a longitudinal compression due to the tension of the tie-rods, and hence to a certain extent resists as a column, the plane of whose flexure is vertical, (since we shall here suppose the beam supported laterally.) Taking the case of uniform loading, (total load = $P$) and supposing the tie-rods screwed up (by sleeve nuts) until the top of the post is on a level with the piers, we know that the pressure between the post and the beam is $P' = \frac{5}{4}W$ (see § 275). Hence by the parallelogram of forces (see Fig. 322) the tension in each tie-rod is $$Q = \frac{P'}{2\cos\alpha} = \frac{5}{16}\frac{W}{\cos\alpha}$$ At each pier the horizontal component of $Q$ is $$P - Q\sin\alpha = \frac{5}{16}W\tan\alpha$$ (1) Hence we are to consider the half-beam $BO$ as a "pin-and-square" column under a compressive force $P' = \frac{5}{4}W\tan\alpha$, as well as a portion of a continuous vertical girder over three equilibrium sections, under a vertical load $W$. The beam carries uniform load $W$. In the outer flange of the dangerous section, $O$, (see also § 273 and Fig. 275) the compression per sq. inch due to both these straining actions must not exceed a safe limit, $P'$, (see § 251). In eq. (6) § 307, where $P'$ is the breaking force for a pin-and-square column, the great- 382 MECHANICS OF ENGINEERING. est stress in any outer fibre = $C$ (the Modulus of Crush- ing) per unit of area. If then we write $p_{\text{out}}$ instead of $C$ in that equation, and $\frac{1}{4}W$ $T$tan a instead of $P$, we have $$\max \text{stress due to column section} = p_{\text{out}} - \frac{5}{16} \cdot \frac{W\tan a}{F}\left[1 + \frac{16}{F} \beta^2 D^2\right];$$ while from eq. (3), p. 396, we have (remembering that our present $W$ represents double the $W$ of § 273), $$\max \text{stress due to girder action} = p_{\text{out}} - \frac{1}{16} \cdot \frac{Wl}{F}\left[\frac{1}{l} + \frac{1}{F}\right].$$ By writing $p_{\text{out}} + p_{\text{in}} = R$ a safe value of compression per unit-area, we have the equation for safe loading $$W\left[5\tan a(1 + \frac{16}{F}\beta^2) + \frac{le}{F}\right] = 16FR.\quad(3)$$ Here $l$ is the half-span $OB$, Fig. 322, $e$ is the distance of outer fibre from the horizontal gravity axis of the cross-section, $k^2$ is the radius of gyration of the section referred to the same axis, while $F$ = area of section. $\beta$ should be taken from the end of § 307. EXAMPLE. — If the span is 30 ft. = 900 in., the girder a 15 inch heavy I-beam of wrought iron, 200 lbs. to the yard, in which case $a$ is of the order of 7/8 inches, $l$ = 90 sq. in., and $k$ = 35.3 sq. inches (taken from the Trenton Co.'s hand-book), required the safe load $W'$, the strut being 5 ft. long. From § 307, $\rho = 1: 36,000$; tan $a = 1 + 5/8 = 3/4$. Hence, using the units pound and inch throughout, and putting $l' = 12,000$ lbs. per sq. in. = max allowable compression stress, we have from eq. (2) $$W' = \frac{16 \times 30 \times 12,000}{15\left[1 + \frac{16}{F}\beta^2(180) + \frac{180 \times k^2}{F}\right]} = 71,111 \text{ lbs}.$$ FLEXURE LONG COLUMNS. 383 only 11.5 tons. [Let the student design the tie-rods (and the strut)]. 314. Buckling of Web-Plates in Built Girders—in §257 mention was made of the fact that very high web plates in built beams, such as I-beams and box-girders, might need to be stiffened by riveting T-irons on the sides of the web. (The girders here spoken of are horizontal ones, such as might be used for carrying a railroad over a short span of 20 to 30 feet.) An approximate method of determining whether such stiffening is needed to prevent lateral buckling of the web, may be based upon Rankine's formula for a long column and will now be given. Fig. 322 shows, free, a portion of a bent I-beam, between two vertical sections at a distance apart—h—the height of the web. In such a beam under forces L to its axis it has been proved (§256) that we may consider the web to sustain all the shear, J, at any section, and the flanges to take all the tension and compression, which form the "stress-couple" of the section. These couples and the two shears are shown in Fig. 323, for the two exposed sections. There is supposed to be no load on this portion of the beam, hence the shears at the two ends are A diagram showing a bent I-beam with two vertical sections labeled A and B, separated by a distance h equal to the height of the web. A diagram showing a bent I-beam with two vertical sections labeled A and B, separated by a distance h equal to the height of the web. A diagram showing a bent I-beam with two vertical sections labeled A and B, separated by a distance h equal to the height of the web. equal. Now the shear acting between each flange and the horizontal edge of the web is equal in intensity per square inch to that in the vertical edge of the web; hence if the web alone, of Fig. 323, is shown as a free body in Fig. 324, we must insert two horizontal forces — J, in opposite 384 MECHANICS OF ENGINEERING. directions, on its upper and lower edges. Each of these $-J$ since we have taken a horizontal length $h_1$ - height of web. In this figure, 324, we notice that the effect of the acting forces is to lengthen the diagonal $BD$ and shorten the diagonal $AC$, both of those diagonals making an angle of 45° with the horizontal. Let us now consider the shearing tendency along $AC$, by treating as free the strip $AC$, of small width = $b$. This is shown in Fig. 325. The only forces acting in the direction of its length $AC$ are the components along $AC$ of the four forces $J'$ at the extremities. We may therefore treat the strip as long column of length $l$, $\sqrt{2}l$, of a section area $F = b\delta_0$, (where $b$ is the thickness of the web plate), with a value of $k' = \frac{1}{\sqrt{2}}b$ (see 309), and with fixed (or flat) ends. Now the sum of the longitudinal components of the two $J'$'s at A is $Q = 2 J'\sqrt{2}$ = $J'\sqrt{2}$; but $J'$ itself = $\frac{b}{\sqrt{2}}$, $\frac{b}{\sqrt{2}}$, $\frac{b}{\sqrt{2}}$, since the small rectangle on which $J'$ acts has an area = $\frac{b}{\sqrt{2}} \cdot \frac{b}{\sqrt{2}}$, and the shearing stress on it has an intensity of $(J + h)$ per unit of area. Hence the longitudinal force at each end of this long column is $$Q = \frac{b}{h_1}J \quad \ldots \quad \ldots \quad \ldots \quad (1)$$ According to eq. (4) and the table in § 307, the safe load factor of safety = 4 for a wrought-iron column of this form, with flat ends, would be (pound and inch) $$P_1 = \frac{1}{4}\left(\frac{b}{h_1}\right)\left(36,000\right) = \frac{9,000}{h_1} \quad \ldots \quad (2)$$ $$1 + \frac{1}{\sqrt{2}}h_1 = 1 + \frac{1}{\sqrt{2}}h_1 \quad \ldots \quad (3)$$ $$36,000\sqrt{\frac{1}{h_1^2}} = 1,500\sqrt{\frac{1}{h_1^2}}$$ If, then, in any particular locality of the girder (of wrought-iron) we find that $Q > P_1$, i.e. if $$\frac{\sqrt{2}}{h_1} > \frac{9,000}{h_1}$$ (pound and inch) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . $$\frac{\sqrt{2}}{h_1} > 1 - \frac{1}{h_1}$$ (pound and inch) FLEXURE. LONG COLUMNS. 385 then vertical stiffeners will be required laterally. When these are required, they are generally placed at intervals equal to $h$, (the depth of web), along that part of the girder where $Q$ is $> P_r$. EXAMPLE Fig. 326.—Will stiffening pieces be required in a built girder of 20 feet span, bearing a uniform load of 40 tons, and having a web 24 in. deep and $\frac{1}{4}$ in. thick? From (3) we know that the $J$ max. of the girder shown in Fig. 326 is close to either pier, and hence we investigate that part of the girder first. $$J_{\text{max}} = \frac{1}{2} W = 20 \text{ tons} = 40,000 \text{ lbs.}$$ :. (inch and lb.), see (3), $$\frac{J}{h} = \frac{40,000}{24} = 1666.6 \quad \ldots \quad (4)$$ while, see (3), (inch and pound), $$\frac{9,000 \times \frac{1}{4}}{1 + \frac{1}{500} (\frac{1}{3})^2} = 905.0 \quad \ldots \quad (5)$$ which is less than 1666.66. Hence stiffening pieces will be needed near the extremities of the girder. Also, since the shear for this case of loading diminishes uniformly toward zero at the middle they will be needed from each end up to a distance of $\frac{1}{4}$ of 10 ft. from the middle. A diagram showing a built girder with stiffeners. 386 MECHANICS OF ENGINEERING. CHAPTER VII. LINEAR ARCHES (OF BLOCKWORK). 315. A Blockwork Arch is a structure, spanning an opening or gap, depending, for stability, upon the resistance to compression of its blocks, or voussoirs, the material of which, such as stone or brick, is not suitable for sustaining a tensile strain. Above the voussoirs is usually placed a load of some character, (e.g. a roadway,) whose pressure causes the voussoirs to be stressed vertically only. This condition is not fully realized in practice, unless the load is of cut stone, with vertical and horizontal joints resting upon voussoirs of corresponding shape (see Fig. 327), but sufficiently so to warrant its assumption in theory. Symmetry of form about a vertical axis will also be assumed in the following treatment. 316. Linear Arches.—For purposes of theoretical discussion the voussoirs of Fig. 327 may be considered to become infinitely small and infinite in number, thus forming a "linear arch," while retaining the same shapes, their depth $\frac{1}{T}$ to the face being assumed constant that it may not appear in the formulas. The joints between them are $\frac{1}{T}$ to the curve of the arch, i.e., adjacent voussoirs can exert pressure on each other only in the direction of the tangent-line to that curve. A diagram showing a linear arch made up of voussoirs. LINEAR ARCHES. 387 S17. Inverted Catenary, or Linear Arch Sustaining its Own Weight Alone.—Suppose the infinitely small voussoirs to have weight, uniformly distributed along the curve, weighing $q$ lbs. per running linear unit. The equilibrium of such a structure, Fig. 328, is of course unstable but theoretically possible. Required the form of the curve when equilibrium exists. The conditions of equilibrium are, firstly, that the resultant of the vertical forces acting between any two adjacent voussoirs at any point, $A$, of the curve must be tangent to the curve; and secondly, considering a portion $BA$ as a free body, the resultant of $H$, the pressure at $B$ the crown, and $T$ at $A$, must balance $R$ the resultant of the vertical forces (i.e., weights of the elementary voussoirs) acting between $B$ and $A$. But the conditions of equilibrium of a flexible, inextensible and uniformly loaded cord or chain are the very same (weights uniform along the curve) the forces being reversed in direction. Fig. 329. Instead of compression we have tension, while the vertical forces act toward instead of away from, the axis $X$. Hence the curve of equilibrium of Fig. 328 is an inverted catenary (see § 48) whose equation is $$y + c = \frac{1}{2} e \left[ \frac{x^2}{e} - \frac{z^2}{e} \right] \quad \ldots \ldots \quad (1)$$ See Fig. 330. $e = 2.71898$ the Napierian Base. The "parameter" $x$ may be determined by putting $z = a$, the half span, and $y = OY$, the rise, then solving for $c$ by successive Description: A diagram showing a curved arch with various labeled points and lines. Description: A diagram showing a curved arch with various labeled points and lines. Description: A diagram showing a curved arch with various labeled points and lines. 388 MECHANICS OF ENGINEERING. approximations. The "horizontal thrust," or $H_x$ is $-q\alpha$, while if $s$ = length of arch $OA$, along the curve, the thrust $T$ at any point $A$ is $$T = \sqrt{H_x^2 + q^2s^2} \quad \ldots \quad (2)$$ From the foregoing it may be inferred that a series of voussoirs of finite dimensions, arranged so as to contain the catenary curve, with joints $\gamma$ to that curve and of equal weights for each segment of arc will be in equilibrium, and moreover in stable equilibrium on account of friction, and the finite width of the joints; see Fig. 331. 318. Linear Arches under Given Loading.—The linear arches to be considered further will be treated as without weight themselves but as bearing vertically pressing loads (each voussoir being assumed to bear its own weight). Problem.—Given the form of the linear arch itself, it is required to find the law of vertical depth of loading under which the given linear arch will be in equilibrium. Fig. 332, given the curve $ABC$, i.e., the linear arch itself, required the form of the curve $MON$, or upper limit of loading, such that the linear arch $ABC$ shall be in equilibrium under the loads lying between the two curves. The loading is supposed homogeneous and of constant depth $\gamma$ to paper; so that the ordinates between the two curves are proportional to the load per horizontal line at each. Assume a height of load $z$, at the crown, at pleasure; then required the $s$ of any point $A$ as a function of $z$, and the curve $ABC$. A diagram showing a linear arch with various points labeled A, B, C, D, E, F, G, H, I, J. Fig. 331. A diagram showing a linear arch with various points labeled A, B, C, D, E, F, G, H, I, J. Fig. 332. LINEAR ARCHES. 389 **Practical Solution.—Since a linear arch under vertical pressures is nothing more than the inversion of the curve assumed by a cord loaded in the same way, this problem might be solved mechanically by experimenting with a light cord, Fig. 333, to which are hung other heavy cords, or bars of uniform weight per unit length, and at equal horizontal distances apart when in equilibrium. By varying the lengths of the bars, and their points of attachment, we may finally find the curve sought, $M\!N$. (See also § 343.) **Analytical Solution.—Consider the structure in Fig. 334. A number of rods of finite length, in the same plane, are in equilibrium, bearing the weights $P$, $P_1$, etc., at the connecting joints, each piece exerting a thrust $T$ against the adjacent joint. The joint $A$, (the "pin" of the hinge), imagined separated from the contiguous rods and hence free, is held in equilibrium by the vertical force $P$ (a load) and the two thrusts $T$ and $T'$, making angles $\theta$ and $\theta'$ with the vertical; Fig. 335 shows the joint $A$ free. From $Z$ (horizontal comp.) = 0, we have $$T \sin \theta = T' \sin \theta'.$$ That is, the horizontal component of the thrust in any rod is the same for all ; call it $H$. : $$T = \frac{H}{\sin \theta}$$ (1) Fig. 384. Fig. 385. 390 MECHANICS OF ENGINEERING. Now draw a line $ds$ to $T$ and write $F$ (ompona. 1 to $ds$) = 0; whence $P \sin \theta = T \sin \beta$, and [see (1)] $\therefore P = \frac{H_c \sin \beta}{\sin \theta \sin \theta'}$ (2) Let the rods of Fig. 334 become infinitely small and infinite in number and the load continuous. The length of each rod becomes $= ds$ an element of the linear arch. $\beta$ is the angle between two consecutive $ds$, $\theta$ is the angle between the tangent line and the vertical, while $P$ becomes the load resting on a single $ds$, or horizontal distance between the middle of the two $ds$. That is, Fig. 336, if $r$ = weight per unit length of arch, then loading $P = r ds$. (The lamina of arch and load considered is unity.) $T$ to paper, in thickness.) $H_c$ = constant = thrust at crown $O$; $\theta = \theta'$, and $\sin \beta = d - p_o$, (since the angle between two consecutive tangents is that between two consecutive radii of curvature). Hence eq. (2) becomes $$rdx = \frac{H_c ds}{p \sin^2 \theta}$$ but $dx = ds \sin \theta$, $\therefore yz = \frac{H_c}{p \sin^2 \theta}$ (3) Call the radius of curvature at the crown $p_o$ and since there $s = x_0$ and $\theta = 90^\circ$, (3) gives $y p_o = H_c$; hence (3) may be written $$s = \frac{H_c}{p \sin^2 \theta}$$ (4) This is the law of vertical depth of loading required. For a point of the linear arch where the tangent line is vertical, $\sin \theta = 0$ and $s$ would = $\infty$ ; i.e., the load would be in- LINEAR ARCHES. 391 **Fretfully high.** Hence, in practice, a full semi-circle, for in- stance, could not be used as a linear arch. 319. Circular Are as Linear Arch.—As an example of the preceding problem let us ap- ply eq. (4) to a circular arc, Fig. 337, as a linear arch. Since for a circle $\rho$ is con- stant $=r$, eq. (4) reduces to $$z = \frac{r}{\sin \theta}$$ (5) Fig. 337. Hence the depth of loading must vary inversely as the cube of the sine of the angle $\theta$ made by the tangent line (of the linear arch) with the ver- tical. To find the depth $z$ by construction.—Having $x$, given, $O$ being the centre of the arch, prolong $Ca$ and make $ab = x$; at $b$ draw a $\gamma$ to $Cb$, intersecting the vertical through $a$ at some point $d$; draw the horizontal do to meet $Cb$ at some point $c$. Again, draw ce to $Cc$, meeting ad in $e$; then ae = x required; a being any point of the linear arch. For, from the similar right triangles involved, we have $$z = ab - ac \sin \theta = ac \sin \theta.$$ sin $\theta - ae \sin \theta = ae \sin \theta$$ $$z = \frac{ae}{\sin \theta}; i.e., ae = z \cdot \frac{\sin \theta}{\sin \theta}$$ Q.E.D. (see (5)) 320. Parabola as Linear Arch.—To apply eq. 4 § 318 to a parabola (axis vertical) as linear arch, we must find values of $\rho$ and $\rho_a$ the radii of curvature at any point and the crown respectively. That is, in the general formula, $$\rho = \left[1 + \left(\frac{dy}{dx}\right)^2\right]^{\frac{1}{2}} + \left(\frac{dy}{dx}\right)^2$$ we must substitute the forms for the first and second dif- ferential co-efficients, derived from the equation of the 392 MECHANICS OF ENGINEERING. Fig. 338. Fig. 339. curve (parabola) in Fig. 338, i.e. from $x^2 = 2py$; whence we obtain $$\frac{dy}{dx} \text{or } \cot \theta = \frac{x}{p} \text{ and } \frac{dy}{dx} = \frac{1}{p}$$ Hence $\rho = \sqrt{\frac{1 + \cot^2 \theta}{1 + p}} = p \cos \theta$, i.e. $\rho = \frac{p}{\sin \theta}$ . . . . (6) At the vertex $\theta = 90^\circ$; $\rho_0 = p$. Hence by substituting for $\rho$ and $\rho_0$ in eq. (4) of § 318 we obtain $$z = z_0 = \text{constant}$$ [Fig. 339] . . . . (7) for a parabolic linear arch. Therefore the depth of homo- geneous loading must be the same at all points as at the corner; i.e., the load is uniformly distributed with respect to the horizontal. This result may have been anis- cipated from the fact that it can be assumed the parabolic form when its load (as approximately true for suspension bridges) is uniformly distributed horizontally. See § 46 in Statics and Dynamics. 321. Linear Arch for a Given Upper Contour of Loading, the arch itself being the unknown lower contour. Given the upper curve or line of load and the depth $z_0$ at corner, required the form of the arch, so that it may be in equilib- rium under the homogeneous load between it and that upper curve. In Fig. 340 let $MON$ be the given upper contour of load, $z_0$ is given or assumed, $x'$ and $y'$ are the respective ordinates of the two curves $BAC$ and $MON$. Required the equation of $BAC$. $$\frac{dy}{dx} = \frac{1}{p}$$ LINEAR ARCHES 393 Fig. 548. Fig. 549. As before, the loading is homogeneous, so that the weights of any portions of it are proportional to the corresponding areas between the curves. (Unity thickness $\mathfrak{I}$ to paper.) Now, Fig. 341, regard two consecutive $d\ell$'s of the linear arch as two links or consecutive blocks bearing at their junction $m$ the load $dP = \gamma (\ell + x') dx$ in which $\gamma$ denotes the heaviness of weight of a cubic unit of the loading. If $T$ and $T'$ are the thrusts exerted on these two blocks by their neighbors (here supposed removed) we have the three forces $dP$, $T$ and $T'$, forming a system in equilibrium. Hence from $2X = 0$, $$T \cos \varphi = T' \cos \varphi' \quad \ldots \quad \ldots \quad (1)$$ and $$\Sigma Y = 0 \text{ gives } T \sin \varphi' - T' \sin \varphi = dP \quad \ldots \quad (2)$$ From (1) it appears that $T \cos \varphi$ is constant at all points of the linear arch (just as we found in § 318) and hence — the thrust at the crown, — $H$, whence we may write $$T = H + \cos \varphi' \quad T' = H - \cos \varphi' \quad \ldots \quad (3)$$ Substituting from (3) in (2) we obtain $$H (\tan \varphi' - \tan \varphi) = dP \quad \ldots \quad (4)$$ But $\tan \varphi' = \frac{dx'}{dx}$ and $\tan \varphi' = \frac{dx + dx'}{dx}$, ($dx$ constant) while $dP = \gamma (\ell + x') dx$. Hence, putting for convenience $H = 7a'$, where $a'$ is side of an imaginary square of the 394 MECHANICS OF ENGINEERING. loading, whose thickness = unity and whose weight = $H$) we have. $$\frac{dx'}{dx} = \frac{1}{a^2}(x' + x'') \quad \ldots \ldots \quad (5)$$ as a relation holding good for any point of the linear arch which is to be in equilibrium under the load included between itself and the given curve whose ordinates are $x'$, Fig. 340. 222. Example of Proceeding. Upper Contour a Straight Line—Fig. 342. Let the upper contour be a right line and horizontal; then the $x''$ of eq. 5 becomes zero at all points of $ON$. Hence drop the ascent of $x'$ in eq. (5) and we have $$\frac{dx}{dx'} = \frac{z}{a^2} \quad \ldots \ldots \quad (6)$$ Multiplying which by $dx$ we obtain $$\frac{dz}{dx'} = \frac{1}{a^2} dz \quad \ldots \ldots \quad (6)$$ This being true of the $z$, $dx$, $dx'$ and $dz$ of each element of the curve $OZ$ whose equation is desired, conceive it written out for each element between $O'$ and any point $m$, and put the sum of the left-hand members of these equations — to that of the right-hand members, remembering that $a^2$ and $dz^2$ are the same for each element. This gives $$\frac{1}{a^2}\int_{m}^{m'} dx dx' - \frac{1}{a^2}\int_{m}^{m'} dz; \text{i.e., } \frac{1}{a^2}\left(\frac{x'}{z'} - \frac{d}{dz}\right)$$ $$\therefore dz = \sqrt{\frac{ax}{z'}} - a\cdot\sqrt{\frac{(z')}{z}} - 1 \quad \ldots \ldots (7).$$ LINEAR ARCHES. 395 Fig. 342 Fig. 343 Integrating (7.) between $O'$ and any point $m$ $$\frac{z}{x} = a \left[ \log_{e} \left( \frac{z}{x} + \sqrt{\left( \frac{z}{x} \right)^2 - 1} \right) \right] \dots (8')$$ i.e., $x = a \log_{e} \left[ \frac{z}{x} + \sqrt{\left( \frac{z}{x} \right)^2 - 1} \right]; \dots (8)$ or $z = \frac{x}{a} \left[ e^x + e^{-x} \right] \dots (9)$ This curve is called the transformed catenary since we may obtain it from a common catenary by altering all the ordinates of the latter in a constant ratio, just as an ellipse may be obtained from a circle. If in eq. (9) $a$ were $= z$, the curve would be a common catenary. Suppose $z_0$ and the co-ordinates $x_0$ and $y_0$ of the point $B$ (abutment), and let us compute $a$ from eq. 8 by putting $x = x_0$, and $z = z_0$, and solving for $a$. Then the crown-thrust $H = ya^2$ becomes known, and $a$ can be used in eqs. (8) or (9) to plot points in the curve or linear arch. From eq. (9) we have $$\text{area } I = \int_{x_0}^{z_0} dx = \frac{z_0}{2} \int_{x_0}^{z_0} \left[ e^x - e^{-x} \right] dx = \frac{as}{2} \left[ e^x - e^{-x} \right] . . . (10)$$ Fig. 343. Call this area, $A$. As for the thrust at the different joints of the linear arch, see Fig. 343, we have crown-thrust $H = ya^2$ ; . . . (11) and at any joint $m$ the thrust $$T = \sqrt{H^2 + (yA)^2} = r\sqrt{a^2 + d^2} . . . (12)$$ 396 MECHANICS OF ENGINEERING. 322. Remarks.—The foregoing results may be utilised with arches of finite dimensions by making the arch-rings contain the imaginary linear arch, and the joints $J$ to the curve of the same. Questions of friction and the resistance of the material of the voussoirs are reserved for succeeding chapter, (§ 344) in which will be advanced a more practical theory dealing with approximate linear arches or "equilibrium polygons" as they will then be called. Still, a study of exact linear arches is valuable on many accounts. By inverting the linear arches so far presented we have the forms assumed by flexible and inextensible cords loaded in the same way. A torn page from a book. GRAPHICAL STATICS. 397 CHAPTER VIII ELEMENTS OF GRAPHICAL STATICS. 284. Definition.-In many respects graphical processes have advantages over the purely analytical, which recommend their use in many problems where celerity is desired without refined accuracy. One of these advantages is that gross errors are more easily detected, and another that the relations of the forces, distances, etc., are made so apparent to the eye, in the drawing, that the general effect of a given change in the data can readily be predicted at a glance. Graphical Statics is the system of geometrical construe- tions by which problems in Statics may be solved by the use of drafting instruments, forces as well as distances being represented in amount and direction by lines on the paper, of proper length and position, according to arbitrary scales; so many feet of distance to the linear inch of paper, for example, for distances; and so many pounds or tons to the linear inch of paper for forces. Of course results should be interpreted by the same scale as that used for the data. The parallelogram of forces is the basis of all constructions for combining and resolving forces. 285. Force Polygons and Concurrent Forces in a Plane.—If a material point is in equilibrium under three forces $P_1$, $P_2$, $P_3$ (in the same plane of course) Fig. 344, any one of them, A diagram showing three concurrent forces acting on a point in equilibrium. 398 MECHANICS OF ENGINEERING. as $P_1$ must be equal and opposite to $R$ the resultant of the other two (diagonal of their parallelogram). If now we lay off to some convenient scale a line in Fig. 345 = $P_1$ and 1 to $P_1$ in Fig. 346, we shall find that one end of $P_1$ is line equal and 1 to $P_1$, and laid off pointing the same way, we note that the line remaining to close the triangle in Fig. 345 must be = and 1 to $P_2$, since that triangle is nothing more than the left-hand half-parallelogram of Fig. 344. Also, in 345, to do the triangle properly, the continuous Point to Butt, round the periphery, Fig. 345 is called a force polygon; of three sides only in this case. By means of it, given any two of the three forces which hold the point in equilibrium, the third can be found, being equal and 1 to the side necessary to "close" the force polygon. Similarly, if a number of forces in a plane hold a material point in equilibrium, Fig. 346, their force polygon, A diagram showing a force polygon with points labeled A, B, C, D, E. Fig. 347, must close, whatever be the order in which its sides are drawn. For, if we combine $P_1$ and $P_2$ into a resultant $O_1$, Fig. 346, then this resultant with $P_3$ to form a resultant $O_2$, and so on; we find the resultant of $P_1$, $P_2$, $P_3$ and $P_4$ to be $O_3$, and if a fifth force is to produce equilibrium it must be equal and opposite to $O_3$, and would close the polygon (side) $O_4$, in which the sides are equal and par- GRAPHICAL STATICS. 399 allel respectively to the forces mentioned. To utilize this fact we can dispense with all parts of the parallelograms in Fig. 346 except the sides mentioned, and then proceed as follows in Fig. 347 : If $P_1$ is the unknown force which is to balance the other four (i.e., their anti-resultant), we draw the sides of the force polygon from $A$ round to $B$, making each line parallel and equal to the proper force and pointing the same way ; then the line $BA$ represents the required $P_2$ in amount and direction, since the arrow $BA$ must follow the course of the arrow $AB$. If the arrow $BA$ were pointed at the extremity $B$, then it gives, obviously, the amount and direction of the resultant of the four forces $P_1,\ldots,P_r$. The foregoing shows that if a system of Concurrent Forces in a Plane is in equilibrium, its force polygon must close. 326. Non-Concurrent Forces in a Plane.—Given a system of non-concurrent forces in a plane, acting on a rigid body, required graphic means of finding their resultant and anti-resultant; also of expressing conditions of equilibrium. The resultant must be found in amount and direction; and also in position (i.e., its line of action must be determined). E. g., Fig. 348 shows a curved rigid beam fixed in a vise at $T_1$ and also under the action of forces $P_1,\ldots,P_r$ ($bendes$ the action of the vise); required the resultant of $P_1,\ldots,P_r$ and $P_r$. By the ordinary parallelogram of forces we combine $P_1$ and $P_r$ at $a$, the intersection of their lines of action; then $R_1$, with $P_r$ at $b$, to form $R_2$, and finally $R_2$ with $P_r$ at $c$ to form $R_3$, which is : -the resultant required, i.e., of $P_1,\ldots,P_r$; and $\ldots\ldots F$ is its line of action. A diagram showing a curved rigid beam fixed in a vise at T1 and acted upon by forces P1, P2, ..., Pr. The lines of action are shown as arrows. Illustration showing a curved rigid beam fixed in a vise at T1 and acted upon by forces P1, P2, ..., Pr. The lines of action are shown as arrows. MECHANICS OF ENGINEERING. The separate force triangles (half-parallellograms) by which the successive partial resultants $R_1$, etc., were found, are again drawn in Fig. 349. Now since $R_a$ acting in the line $...F$, Fig. 348, is the resultant of $P_1...P_n$, it is plain that a force $R'_a$ equal to $R_a$ and acting along $...F$, but in the opposite direction, would balance the system $P_1...P_n$ (the resultant). That is, the forces $P_1P_2P_3$ and $R'_a$ would form a system in equilibrium. The force $R'_a$ then, represents the action of the vice $V$ upon the beam. Hence replace the vice by the force $R'_a$ acting in the line $...F...c$; to do which requires us to imagine a rigid prolongation of that end of the beam, to intersect $...F...c$. This is shown in Fig. 350 where the whole beam is free, in equilibrium, under the forces shown, and in precisely the same state of stress, part for part, as before. In this figure we have one force diagram, in Fig. 351, all the force triangles of Fig. 349 (by making their common sides coincide, and putting $R'_a$ instead of $R_a$, and dotting all forces other than those of Fig. 350), we have a figure to be interpreted in connection with Fig. 350. A diagram showing a half-parallellogram representing a force triangle. Fig. 349. Here we note, first, that in the figure called a force-diagram, $P_1P_2P_3P_4$ and $R'_a$ form a closed polygon and that A diagram showing a space diagram and a force diagram. Fig. 350. Fig. 351. GRAPHICAL STATICS. 407 their arrows follow a continuous order, point to butt, around the perimeter; which proves that one condition of equilibrium of a system of non-concurrent forces in a plane is that its force polygon must close. Secondly, note that ab is to $Oa'$, and bc to $Ob'$; hence if the force-diagram has been drawn (including the rays dotted) in order to determine the amount and direction of $R_1$, or any other one force, we may then find its line of action in the space-diagram, as follows: (N. B.—By space diagram is meant the figure showing to a true scale the form of the rigid body and the lines of action of all the forces.) Draw through $P_1$ and $P_2$ a line to $Oa'$ to cut $P_3$ in some point $b$; then through $b$ a line $l$ to $Ob'$ to cut $P_4$ at some point $c$. $R_1$ drawn to $Oc'$ is the required line of action of $R_1$, the resultant of $P_1$, $P_2$, $P_3$, and $P_4$. abc is called an equilibrium polygon; this one having but two segments, ab and bc (sometimes the lines of action of $F_1$ and $R_1$ may conveniently be considered as segments). The segments of the equilibrium polygon are parallel to the respective rays of the force-diagram. Hence for the equilibrium of a system of non-concurrent forces in a plane not only must its force polygon close, but also the first and last segments of the corresponding equilibrium polygon must coincide with the resultants of the first two forces, and of the last two forces, respectively, of the system. E.g., ab coincides with the line of action of the resultant of $P_1$ and $P_2$; bc with that of $P_3$ and $R_1$. Evidently the equil. polygon will be different for each choice of pole, O. Hence for the force polygons on different choice of a pole, O. But if the order of forces be taken as above, as they occur along the beam, or structure, and the pole taken at the "butt" of the first force in the force polygon, there will be only one (and this one will be called the special equilibrium polygon in the chapter on arch-ribs, and the "true linear arch" in dealing with the stone arch.) After the rays (dotted in Fig. 351) have been added, by joining the pole to each 402 MECHANICS OF ENGINEERING. vertex with which it is not already connected, the final figure may be called the force diagram. It may sometimes be convenient to give the name of rays to the two forces of the force polygon which meet at the pole, in which case the first and last segments of the corresponding equil. polygon will coincide with the lines of action of those forces in the space-diagram (as we may call the representation of the body or structure on which the forces act). This "space diagram" shows the real field of action of the forces, while the force diagram, which is drawn on a plane perpendicular to that of the paper, shows the magnitudes and directions of the forces acting in the former diagram, its lines being interpreted on a scale of so many lbs. or tons to the inch of paper ; in the space-diagram we deal with a scale of so many feet to the inch of paper. We have found, then, that if any vertex or corner of the closed force polygon be taken as a pole, and rays drawn from it to all the other corners of the polygon, and a corresponding ray drawn from each point on these rays to the first and last segments of the latter polygon must co-incide with the first and last forces according to the order adopted (or with the resultant of the first two and last two, if more convenient to classify them thus). It remains to utilize this principle. 327. To Find the Resistant of Several Forces in a Plane.—This might be done as in § 326, but since frequently a given set of forces are parallel, or nearly so, a special method will now be given, of great convenience in such cases. Fig. 352. Let $P_1$, $P_2$ and $P_3$ be the given forces whose resultant is required. Let us first find their one- resultant, or force which will balance Fig. 352. Fig. 352. GRAPHICAL STATICS. 403 them. This anti-resultant may be conceived as decomposed into two components $P$ and $P'$ of one of which, say $P$, is arbitrary in amount and position. Assuming $P$, then, at convenience, in the space diagram, it is required to find $P'$. The five forces must form a balanced system; hence if beginning at $O_0$, Fig. 353, we lay off a line $O_0A = P$ by scale, then $A1 = 1$ to $F_1$, and so on (point to butt), the line $BO_1$ necessary to close the force polygon is $= P'$ required. Now form the corresponding equil. polygon in the space diagram by drawing a line through $b$ o the intersection of $P$ and $F_1$, draw $ab$ to the ray $O_1 - 1$ (which connects the pole $O_1$ with the point of the lastforce mentioned). From $b$, where $ab$ intersects the line of $F_1$, draw $bc$, to the ray $O_1 - 2$, till it intersects the line of $F_2$. A line $me$ drawn through $c$, till it to the $P'$ of the force diagram in the line of action of $P'$. Now the resultant of $P$ and $P'$ is the anti-resultant of $F_1$, $F_2$ and $F_3$... $d$, the intersection of the lines of $P$ and $P'$, is a point in the line of action of the anti-resultant re- quired, while its direction is indicated by the line $BA$ in the force diagram; for BA forms a closed poly- gon both with $P$, $P_1$, $P_2$ and with $PP'$. Hence a line through d if to BA, viz., de, is the line of action of the anti-resultant (and hence of the resultant) of $F_n$, $F_{n+1}$, $F_{n+2}$. Since, in this construction, $P$ is arbitrary, we may first choose $O_0$, arbitrarily, in a convenient position, i.e., at a position that by inspection the segments of the result- ing equil. polygon shall give fair intersections and not pass off at any point. But when they are parallel the device of introducing the oblique $P$ and $P'$ is quite neces- sary. 328.—The result of this construction may be stated as follows, (regarding Oo and cm as segments of the equil. polygon as well as ab and bc): If any two segments of an equil. polygon be prolonged, their intersection is a point in the line of action of the resultant) of those forces acting at 404 MECHANICS OF ENGINEERING the vertices intervening between the given segments. Here, the resultant of $P_1, P_2, P_3$ acts through $d$. 329. Vertical Reaction of Pier, etc. Fig. 354. Given the vertical forces or loads $P_1, P_2$, and $P_3$ acting on a rigid body (beam, or truss) which is supported by two piers having smooth horizontal surfaces (so that the reactions must be vertical), required the reactions $V_a$ and $V_b$ of the piers. For an instant suppose $V_a$ and $V_b$ known; they are in A diagram showing a beam with three vertical forces acting on it, and two piers supporting it. The diagram also shows the force polygon formed by these forces. Fig. 354. equil. with $P_1, P_2$, and $P_3$. The introduction of the equal and opposite forces $P'$ and $P''$ in the same line will not disturb the equilibrium. Taking the seven forces in the order $P, P', P_1, P_2, P_3$, and $P''$, a force polygon formed with them will close (see $(b)$ in Fig., where the forces which really lie on the same line are slightly separated). With $O$, the butt of $P$, as a pole, draw the rays of the force diagram $O d$, $O b$, etc. The corresponding equil. polygon begins at $a$, the intersection of $P$ and $V_a$ in $(a)$ (the space diagram), and ends at $a$ the intersection of $P'$ and $V'_a$ in $(b)$ (the space diagram). Since $P$ and $P'$ act in the same plane, this line must be that line and must be $a$ to $P$ and $P'$ of the force diagram. Since the amount and direction of $P$ and $P'$ are arbitrary, the position of the pole $O$ is arbitrary, while $P_a, P_b$, and $P_c$ are the only forces known in advance in the force diagram. Hence $V_a$ and $V_b$ may be determined as follows: Lay off the given loads $F_1, F_2$, etc., in the order of their occurrence in the space diagram, to form a "load-line" AD GRAPHICAL STATICS. 465 Sec (b) Fig. 354) as a beginning for a force-diagram; take any convenient pole $O$, draw the rays $OA$, $OB$, $OC$ and $OD$. Then beginning at any convenient point $a$ in the vertical line containing the unknown $V_a$, draw ab to $Od$, bc to $Ob$, and so on, until the last segment ($dn$ in this case) cuts the vertical containing the unknown $V_a'$ in some point. Join as (this is sometimes called a closing line) and draw a line through $O$, in the force-diagram. This last line will cut the "load-line" in some point $n'$, and divide it in two parts $a'd$ and $Dn'$, which are respectively Corollary.—Evidently, for a given system of loads, in given vertical lines of action, and for two given piers, or abutments, having smooth horizontal surfaces, the location of the point $n'$ on the load line is independent of the choice of pole. Of course, in treating the stresses and deflection of the rigid body concerned, $P$ and $P'$ are left out of account, as being imaginary and serving only a temporary purpose. 330. Application of Foregoing Principles to a Roof Truss—Fig. 355. $W_1$ and $W_2$ are wind pressures, $P_1$ and $P_2$ are loads, while the remaining external forces, viz., the re- A diagram showing a roof truss with various forces labeled. The diagram includes labels such as "Load Line," "Vertical Lines," "Horizontal Surfaces," and "Load Points." The diagram also includes symbols for different forces like $W_1$, $W_2$, $P_1$, and $P_2$. Fig. 355. 406 MECHANICS OF ENGINEERING. actions, or supporting forces, $V_x$, $V_y$, and $H_x$, may be found by preceding §3. (We here suppose that the right abutment furnishes all the horizontal resistance; none at the left). Lay off the forces (known) $W_x$, $W_y$, $F_x$, and $F_y$ in the usual way, to form a portion of the closed force polygon. To close the polygon it is evident we need only draw a horizontal through 3 and limit it by a vertical through 1. This determines $H_x$, but it remains to determine $n'$ the point of intersection of $V_x$ and $V_y$. Select an arbitrary pole $O_0$, and draw rays from it to 1, 2, etc. Assume a convenient point $a$ in the line of $V_x$ in the space diagram, and through it draw a line to 0,1 to meet the line of $W_x$ in some point b; then a line to $O_2$ to meet the line of $W_y$ in some point c; then through c to 0,3 to meet the line of $P_x$ in some point d; then through d to 0,4 to meet the line of $P_y$ in some point e ($e$ is identical with $d$, since $P_x$ and $P_y$ are in the same line); then q'f1 to 0,5 to meet $H_x$ in some point f; then q'f2 to 0,5 to meet $V_y$ in some point g. $aobcfdg$ is an equilibrium polygon corresponding to the pole $O_0$. Now join ag, the "closing-line," and draw a line to it through $O_1$ to determine $n'$, the required point of intersection between $V_x$ and $V_y$ on the vertical 1-6. Hence $F_x$ and $V_y$ are now determined as well as $H_x$. (The use of the arbitrary pole $O_0$, implies the temporary employment of two equal forces and equal forces in the line ag, the amount of either being = $\theta O_0A$.) Having now all the external forces acting on the truss, and assuming that it contains no "redundant parts," i.e., parts unnecessary for rigidity of the frame-work, we proceed to find the pulls and thrusts in the individual pieces, on the following plan. The truss being pin-connected, no piece extending beyond a joint, and all loads being considered to act at joints, the action pull or thrust, of each piece on the joint at either extremity will be in the direction of the piece, i.e., in a known direction, and the pin of each GRAPHICAL STATICS. 407 joint is in equilibrium under a system of concurrent forces consisting of the loads (if any) at the joint and the pulls or thrusts exerted upon it by the pieces meeting there. Hence we may apply the principles of § 325 to each joint in turn. See Fig. 356. In constructing and interpreting the various force polygons, Mr. R. H. Bow's convenient notation will be used; this is as follows: In the space diagram a capital letter ($AB$, etc.) is placed in each triangular cell of the truss, and also in each angular space in the outside outline of the truss between the external forces and the adjacent truss-pieces. In this way we can speak of the stress in the force $BC$, i.e., $W_1$ as that at $CZ$, the stress in the piece $aB$ as the force $CD$, and so on. That is, the stress in any one piece can be named from the letters in the spaces bordering its two sides. Corresponding to these capital letters in the spaces of the space-diagram, small letters will be used at the vertices of the closed force-polylgons (one polygon for each joint) in such a way that the stress in the piece $CD$, for example, shall be the force of the force polygon belonging to any joint in which $CD$ is a side, and similarly, the stress $FG$ by the force $fg$ in the proper force polygon, and so on. In Fig. 356 the whole truss is shown free, in equilibrium under the external forces. To find the pulls or thrusts (i.e., tensions or compressions) in the pieces, consider that if all but two of the forces of a closed force polygon are known in magnitude and direction, while the directions, only, of those two are known, the whole force polygon may be drawn, thus determining the amounts of those two forces by the lengths of the corresponding sides. We must, : begin with a joint where no more than two pieces meet, as at $a$; (call the joints $a$, $\beta$, $\gamma$, $\delta$, and the corresponding force polygons $a'$, $\beta'$ etc. Fig. 356.) Hence at $a'$ (anywhere on the paper) make ab || and - (by scale) to the known force $AB$ (i.e., $V_1$) pointing at it the upper end, and from this end draw bc = and l to the known force BC (i.e., $W_1$) pointing at it at the lower end. 408 MECHANICS OF ENGINEERING. Diagram showing a polygon with various lines and arrows indicating forces and directions. Fig. 300. To close the polygon draw through c a l to the piece CD, and through a b l to AD; their intersection determines $d$, and the polygon is closed. Since the arrows must point in the same direction, we have a polygon with which the piece CD acts on the pin of the joint $a$ is a force of an amount = ol and in a direction from c toward d; hence the piece CD is in compression; whereas the action of the piece DA upon the pin at $a$ is from d toward $a$ (direction of arrow) and hence DA is in tension. Notice that in constructing the force polygon $a'$ a right-handed (or clock-wise) rotation has been observed in considering in turn the spaces $A B C$ and $D$, round the joint $a$. A similar order will be found convenient in each of the other joints. Knowing now the stress in the piece CD, (as well as in DA) all but two of the forces acting on the pin at the joint $\beta$ are known, and accordingly we begin a force polygon, $\beta'$, for that joint by drawing dc= and l to the dc of polygon $a'$, but pointed in the opposite direction, since the action of CD on the joint $\beta$ is equal and opposite to its action on the joint $a$ (this disregards the weight of the piece). Through c draw ce = and f to the force CE (i.e., $W_2$) and GRAPHICAL STATICS. 409 pointing the same way; then $P_1$ = and $J$ to the load $EF$ (i.e. $P_1$) and pointing downward. Through $f$ draw $a$ to the piece $FG$ and through $d$, $s$ to the piece $GD$, and the polygon is closed, thus determining the stresses in the pieces $FG$ and $GD$. Noting the pointing of the arrows, we readily see that $FO$ is in compression while $GD$ is in tension. Next pass to the joint $\delta$, and construct the polygon $\alpha'$, thus determining the stress $\varphi$ in $GH$ and that $\omega d$ in $AD$; this last force $\omega d$ should check with its equal and opposite $\omega d$ already determined in polygon $\alpha'$. Another check consists in the proper closing of the polygon $\gamma'$, all of whose sides are now known. [A compound stress-diagram may be formed by superposing the polygons already found in such a way as to make them co-inside; but the character of each stress is not so readily perceived then as when they are kept separate.] In a similar manner we may find the stresses in any pin-connected frame-work (in one plane and having no redundant pieces) under given loads, provided all the supporting forces or reactions can be found. In the case of a braced-arch (truss) as shown in Fig. 537, hinged together at both ends and not free to slide laterally upon them, the reactions at $O$ and $B$ depend, in amount and direction, not only upon the equations of Statics, but on the form and elasticity of the arch-truss. Such cases will be treated later under arch-ribs, or curved beams. §32. The Special Equil. Polygon. Its Relation to the Stresses in the Rigid Body.—Reproducing Figs. 350 and 351 in Figs. 538 and 539, (where a rigid curved beam is in equilibrium under the forces $T_1$, $T_2$, $F_1$, $F_2$, and $H_1$,) we call a . . . . . . 410 MECHANICS OF ENGINEERING. the special equil. polygon because it corresponds to a force diagram in which the same order of forces has been observed as that in which they occur along the beam (from left to right here). From the relations between the force A diagram showing a space diagram and a force diagram. The space diagram shows a curved beam with points labeled A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. The force diagram shows these points connected by lines labeled P1, P2, P3, P4, P5, P6, P7, P8, P9, P10, P11, P12, P13, P14, P15, P16, P17, P18, P19, P20. The lines in the space diagram are dotted lines. diagram and equal polygon, this special equil. polygon in the space diagram has the following properties in connection with the corresponding rays (dotted lines) in the force diagram. The stresses in any cross-section of the portion O'A of the beam, are due to $P_1$ alone; those of any cross-section on AB to $P_1$ and $P_n$, i.e., to their resultant $R$, whose magnitude is given by the line OA' in the force diagram, while its line of action is at the first segment of the equil. polygon. Similarly, the stresses in BC are due to $P_2$, $F_2$, and $P_n$, i.e., to their resultant $R_2$, whose magnitude is given by OB' in the force diagram. E.g., if the section at m be exposed, considering O'ABM as a free body, we have (see Fig. 300) the elastic stresses (or internal forces) at m balancing the exterior or "applied forces" $P_2$, $P_3$ and $P_n$. Obviously, then, the stresses at m are just A diagram showing a space diagram and a force diagram. The space diagram shows a curved beam with points labeled A', B', C', D', E', F', G', H', I', J', K', L', M', N', O', P', Q', R', S', T', U', V', W', X', Y', Z'. The force diagram shows these points connected by lines labeled P1', P2', P3', P4', P5', P6', P7', P8', P9', P10', P11', P12', P13', P14', P15', P16', P17', P18', P19', P20'. The lines in the space diagram are dotted lines. A diagram showing a space diagram and a force diagram. The space diagram shows a curved beam with points labeled A", B", C", D", E", F", G", H", I", J", K", L", M", N", O", P", Q", R", S", T", U", V", W", X", Y", Z". The force diagram shows these points connected by lines labeled P1", P2", P3", P4", P5", P6", P7", P8", P9", P10", P11", P12", P13", P14", P15", P16", P17", P18", P19", P20". The lines in the space diagram are dotted lines. GRAPHICAL STATICS. 411 the same as if $R_0$ the resultant of $P_a$, $P_b$ and $P_w$ acted upon an imaginary rigid prolongation of the beam intersecting $bc$ (see Fig. 361). $R_0$ might be called the "anti-stress-resultant" for the portion $BC$ of the beam. We may : state the following: If a rigid body is in equilibrium under a system of Men-Courbure forces in a plane, and the special equilibrium polygon has been drawn, the force diagram is the anti-stress-resultant of that portion of the beam which corresponds to the segment of the equilibrium polygon to which the ray is parallel ; and its line of action is the segment just mentioned. Evidently if the body is not one rigid piece, but composed of a ring of unconnected blocks (or voussoirs), it may be considered rigid only so long as no slipping takes place or disarrangement of the blocks; and this requires that the "anti-stress-resultant" shall lie within the blocks, and that two blocks shall not lie outside the bearing surface of the joint, nor make too small an angle with it, lest tipping or slipping occur. For an example of this see Fig. 362, showing a line of three blocks in equilibrium under five forces. The pressure borne at this point $MN$, is = $R_0$ in the force-diagram and acts in the line ab. The con- straints supposed will therefore force given except one, in amount and position, and that this one could easily be found in amount, as being the side remaining to close the force polygon, while its position would depend on the equil. polygon. But in practice the two forces $P_a$ and $P'_a$ are generally unknown, hence the point $O$, or pole of the force diagram, can not be fixed, nor the special equil. polygon located, until other considerations, outside of those so far presented, are brought into play. In the progress of such a problem as will be seen it will be necessary to draw trial positions for the pole $O$, and corresponding trial equilibrium polygons. Fig. 362. 501 412 MECHANICS OF ENGINEERING. CHAPTER IX. GRAPHICAL STATICS OF VERTICAL FORCES. 338. Remarks.--(With the exception of § 378 a) in problems to be treated subsequently (either the stiff arch-rib, or the block-work of an arching, of masonry) when the body is considered as a whole, the force holding it in equilib will be vertical (loads, due to gravity) except the reactions at the two extremities, as in Fig. 363; but for convenience each reaction will be replaced by its horizontal and vertical components (see Fig. 364). The two $H$'s are of course equal, since they are the only horizontal forces in the system. Henceforth, all equal polygons under discussion will be understood to imply this kind of system of force. $P_{H}$ $P_{V}$ Figs. 365, 366, 367. etc., will represent the "loads"; $V_{a}$ and $V_{b}$ the vertical components of the abutment reactions; $H$ the value of either horizontal component of the same. (We here suppose the pressures $T_{a}$ and $T_{b}$ resolved along the horizontal and vertical.) Fig. 365. Fig. 366. Fig. 367. GRAPHICAL STATICS. 413 334. Concrete Conception of an Equilibrium Polygon.—Any equilibrium polygon has this property, due to its mode of construction, viz.: If the $ab$ and $bc$ of Fig. 358 were imponderable straight rods, jointed at without friction, they would be in equilibrium under the system of forces there given. (See Fig. 364a.) The rod $ab$ suffers a compression equal to the $R_1$ of the force diagram, Fig. 359, and $bc$ a compression = $R_2$. In some cases these rods might be in tension, and would then form a set of links playing the part of a suspension-bridge cable. (See § 44.) 335. Example of Equilibrium Polygon Drawn to Vertical Loads —Fig. 365. [The structure bearing the given loads is not shown, but simply the imaginary rods, or segments of an equilibrium polygon, which would support the given loads in equilibrium if the abutment points $A$ and $B$, to which the terminal rods are hinged, were firm. In the present case this equilibrium is unstable since the rods form a standing structure; but if they were hanging, the equilibrium would be stable. Still, in the present case, a light bracing, or a little friction at all joints would make the equilibrium stable.] A diagram showing three vertical loads $P_1$, $P_2$, and $P_3$ acting on a structure with two abutments A and B. The diagram illustrates how an equilibrium polygon can be drawn to represent the forces acting on the structure. Fig. 365. Given three loads $P_1$, $P_2$, and $P_3$, and two "abutment verticals" $d'$ and $b'$, in which we desire the equil. polygon to terminate, lay off on a "load-line," to scale, $F_1$, $F_2$, and $F_3$ end to end in their order. Then selecting any pole, $$\text{Given three loads } P_1, P_2 \text{ and } P_3 \text{ and two "abutment verticals" } d' \text{ and } b', \text{ in which we desire the equil. polygon to terminate, lay off on a "load-line," to scale, } F_1, F_2 \text{ and } F_3 \text{ end to end in their order. Then selecting any pole,}$$ 414 MECHANICS OF ENGINEERING. $O$, draw the rays $01$, $02$, etc., of a force diagram (the $V$'s and $F$'s, though really on the same vertical, are separated slightly for distinctness); also the $H$'s, which both pass through $O$ and divide the load-line into $V_1$ and $V_2$. We determine a corresponding equilibrium polygon by drawing through $A$ (any point in $A$) a line to $O \cdot 1$, to intersect $P_1$ in some point $b$; through $b$ a line to $O \cdot 2$, and so on until the other abutment-vertical is struck in some point $B$. $AB$ is the "abutment-line" or "closing-line". By choosing another point for $O$, another equilibrium polygon is obtained. There will be an infinite number (which could thus be drawn, for the given loads and the $A$ and $B'$ verticals) is the special equilibrium polygon for the arch-rib or stone-arch, or other structure, on which the loads rest, is to be considered hereafter. In any of the above equilibrium polygons the imaginary series of jointed rods would be in equilibrium. 336. Useful Property of an Equilibrium Polygon for Vertical Loads.-(Particular case of § 328). See Fig. 366. In any equil. polygon, supporting vertical loads, consider as free any number of consecutive segments, or rods, with the loads at their joints, e.g., the 5th and 6th and portions of the 4th and 7th which we suppose cut and the compressive forces in them put in, $T_5$ and $T_7$, in order to consider $4 \cdot 5 \cdot 7$ as free as possible. As in equil., according to Station, the lines of action of $T_5$ and $T_7$ (the compression in these rods) must intersect in a point, $O$, in the line of action of the resultant of $P_5$, $P_6$ and $P_7$; i.e., of the loads occurring at the intervening vertices. That is, the point $C$ must lie in the vertical containing the centre of gravity of those loads. Since the position of this vertical must be independent of the particular equilibrium polygon used, any other dotted lines in Fig. 366 for the same loads will give the same result. Fig. 366. 366 GRAPHICAL STATICS. 415 sults. Hence the vertical $CD$, containing the centre of gravity of any number of consecutive loads, is easily found by drawing the equilibrium polygon corresponding to any convenient force diagram having the proper load-line. This principle can be advantageously applied to finding a gravity-line of any plane figure, by dividing the latter into parallel strips, whose areas may be treated as loads applied in their respective centres of gravity. If the strips are similar, then the line of gravity of each strip may be considered to be at the centre of the line joining the mid-dles of the two long sides, while their areas may be taken as proportional to the lengths of the lines drawn through these centres of gravity parallel to the long sides and limited by the end-curves of the strips. Hence the "load-line" of the force diagram may consist of these lines, or of their halves, or quarters, etc., if more convenient (§ 376). USEFUL RELATIONS BETWEEN FORCE DI- GRAMS AND EQUILIBRIUM POLYGONS. (for vertical loads.) §37. Résumé of Construction.—Fig. 367. Given the loads $P_1$, etc., their verticals, and the two abutment verticals $A'$ and $B'$, in which the abutments are to lie; we lay off a load-line $1 \ldots 4$, take any convenient pole, $O$, for a force-diagram complete with lateral. For a corresponding equilibrium polygon, assume any point $n$ in the vertical $A'$, for an abutment, and draw the successive abutments $A_1$, $2$, etc., respectively parallel to the inclined lines of the force diagram (rays), thus determining finally the abutment $B$, in $B'$, which ($B$) will not in general lie in the horizontal through $A$. Now join $AB$, calling $AB$ the abutment-line, and draw a parallel to it through $O$, thus fixing the point $n'$ on the 416 MECHANICS OF ENGINEERING. Fig. 367. Fig. 368. load-line. This point $s'$, as above determined, is independent of the location of the pole, $O$, (proved in § 329) and divides the load-line into two portions ($P_1' = 1 \ldots n'$, and $P_n' = s' \ldots 4$) which are the vertical pressures which two supports in the verticals $A'$ and $B'$ would sustain if the given loads rested on a horizontal rigid bar, as in Fig. 368. See § 329. Hence to find the point $s'$ we may use any convenient pole $O$. [N.B.--The forces $V_s$ and $V_{s'}$ of Fig. 367 are not identical with $V'$ and $V_{s'}$, but may be obtained by dropping a line from $O$ to the load-line, thus dividing the load-line into two portions which are $V_s$ (upper portion) and $V_{s'}$. However, if $A$ and $B$ be connected by a tie-rod, in Fig. 367, the abutments in that figure will bear vertical pressures only and they will be the same as in Fig. 368, while the tension in the tie-rod will be = $O's'$.] **338. Theorem.—The vertical dimensions of any two equilibrium polygons, drawn to the same loads, load-verticals, and abutment-verticals, are inversely proportional to their H's (or "pole distances")." We here regard an equi. polygon and its abutment-line as a closed figure. Thus, in Fig. 369, we have two force-diagrams (with a common load-line, for convenience) and their corresponding equi. polygons, for the same loads and verticals. From § 327 we know that $OA'$ is 1 to $AB$ and $OA_n'$ is 1 to $AB_n'$. Let CD be any vertical cutting the first segments of the two equi. polygons. 5 GRAPHICAL STATICS. 417 Denote the intercepts thus determined by $z'$ and $z''$ respectively. From the parallels just mentioned, and others more familiar, we have the triangle $O_1z'$ similar to the triangle $A_1z'$ (shaded), and the triangle $O_1z''$ similar to the triangle $A_1z''$. Hence the proportions between $\frac{1}{H} = \frac{z'}{h}$ and $\frac{1}{H} = \frac{z''}{h}$ $\therefore z' : z'' = H : H$. The same kind of proof may easily be applied to the vertical intercepts in any other segments, e.g., $z'''$ and $z''''$. Q. E. D. 339. Corollaries to the foregoing. It is evident that: (1.) If the pole of the force-diagram be moved along a vertical line, the equilibrium polygon changing its form in a corresponding manner, the vertical dimensions of the equilibrium polygon remain unchanged; and (2.) If the pole move along a straight line which contains the point $z'$, the direction of the abutment-line remains constantly parallel to the former line, while the vertical dimensions of the equilibrium polygon change in inverse proportion to the distance of the pole from the force-diagram. ($H$ is the distance of the pole from the load-line, and is called the pole-distance). § 340. Linear Arch as Equilibrium Polygon.—(See § 316.) If the given loads are infinitely small with infinitely small horizontal spaces between them, any equilibrium polygon becomes a linear arch. Graphically we can not deal with these infinitely small loads, but it is evident that if we replace them, in successive groups, A diagram showing a triangular section of a linear arch, with labels A, B, C, D, P1, P2, P3, P4. Fig. 305 The proportions between $\frac{1}{H} = \frac{z'}{h}$ and $\frac{1}{H} = \frac{z''}{h}$ 418 MECHANICS OF ENGINEERING. by finite forces, each of which = the sum of those composing one group, and applied through the centre of gravity of that group, we can draw an equilibrium polygon whose segments will be tangent to the curve of the corresponding line arch, and indicate its position with sufficient exactness for practical purposes. (See Fig. 370.) The successive points of tangency $A$, $m$, etc., lie vertically under the points of contact between the groups. This relation forms the basis of the graphical treatment of voussoir, or blockwork, arches. 341. To Pass an Equilibrium Polygon Through Three Arbitrary Points.—(In the present case the forces are vertical. For a construction dealing with any plane system of forces see construction in § 373.) Given a system of loads $P_1$, $P_2$, etc., outside the load-vertices, the first point being between the verticals of the first two. See Fig. 371. The loads $P_1$, etc., are given, with their vertices, while $A$, $p_1$, and $B$ are the three points. Lay off the load-line, adjoin with the point $O_1$ at $A$. Draw a force $O_1A_1$ then a corresponding preliminary equilibrium polygon beginning at $A$. Its right abutment $B_1$, in the vertical through $B$, is thus found. $O_1a'$ can now be drawn to $A_1B_1$ to determine $a'$. Draw $a'a'$ to $BA$. The pole of the required equilibrium polygon must lie on $a'O$ (§ 37). A diagram showing a line arch with a series of vertical lines labeled A, m, etc., and a polygon labeled P1, P2, etc., tangent to the curve of the corresponding line arch. Fig. 370. A diagram showing a line arch with a series of vertical lines labeled A, m, etc., and a polygon labeled P1, P2, etc., tangent to the curve of the corresponding line arch. Fig. 371. GRAPHICAL STATICS. 419 Draw a vertical through $p$. The $H$ of the required equilibrium polygon must satisfy the proportion $H : H_1 = rs : pm$. (See § 388). Hence construct or compute $H$ from the proportion and draw a vertical at distance $H$ from the load-line (on the left of the load-line here); its intersection with $\pi'$ gives $O$ the desired pole, for which a force diagram may now be drawn. The corresponding equilibrium polygon beginning at the first point $A$ will also pass through $p$ and $B$; it is not drawn in the figure. 348. Symmetrical Case of the Foregoing Problem.--If two points $A$ and $B$ are on a level, the third, $p$, on the middle vertical between them; and the loads (an even number) symmetrically disposed both in position and magnitude, about $p$, we may proceed more simply, as follows: (Fig. 372.) From symmetry $\pi'$ may occur in the middle of the triangle, of which we need lay off only the upper half. Take a convenient pole $O_p$ in the horizontal through $\pi'$, and draw a half force diagram and a corresponding half equilibrium polygon (both dotted). The upper segment $\beta$ of the latter must be horizontal and being prolongation of the prolongation of the first segment in a point $d$, while determining this point $C$ to determine the centre of gravity of the loads occurring over the half-span on the left. (See § 386.) In the required equilibrium polygon the segment containing the point $p$ must be horizontal, and its intersection with the first segment must lie in $CD$. Hence determine this intersection, $C$, by drawing the vertical $CD$ and a horizontal through $p$; then join $Cd$, which is the first segment of the required equil. polygon. A parallel to $Cd$ through 1 is the first ray of the corresponding force diagram, and determines the pole $O$ on the horizontal through $\pi'$. Completing the force diagram for A diagram showing a symmetrical case of equilibrium problem. Fig. 372 420 MECHANICS OF ENGINEERING. this pole (half of it only here), the required equil. poly- gon is easily finished afterwards. 343. To Find a System of Loads Under Which a Given Equi- librium Polygon Would be in Equilibrium—Fig. 373. Let $AB$ be the given equilibrium polygon. Through any point $O$ as a pole draw a parallel to each segment of the equilibrium polygon. Any vertical, as $V$, cutting these lines will have, intercepted upon it, a load which is proportional to $1$, $2$, $3$, etc., are proportional to the successive loads which, placed on the corresponding joints of the equilibrium polygon would be supported by it in equilibrium (unstable). One load may be assumed and the others constructed. A hanging, as well as a standing, equilibrium polygon may be dealt with in like manner, but will be in stable equilibrium. The problem in § 44 may be solved in this way. A diagram showing a line AB with points O and P labeled. A line is drawn from O perpendicular to AB at point P. P. E. ARCHES OF MASONRY. 421 CHAPTER X. RIGHT ARCHES OF MASONRY. 344.—In an ordinary "right" stone-arch (i.e., one in which the faces are $\parallel$ to the axis of the cylindrical soffit, or under surface), the successive blocks forming the arching are called vousoirs, the joints between them being planes which, prolonged, meet generally in one or more horizontal lines; e.g., those of a three-centred arch, three horizontal lines, etc.; those of a segmental arch, the axis of the cylinder, etc. Elliptic arches are sometimes used. The inner concave surface is called the soffit, to which the radiating joints between the vousoirs are made perpendicular. The curved line in which the soffit is intersected by a plane A diagram showing the construction of an arch with vousoirs and intrados. $\parallel$ to the axis of the arch is the Intrados. The curve in the same plane as the intrados, and bounding the outer extremities of the joints between the vousoirs, is called the Extrados. Fig. 374 gives other terms in use in connection with a 422 MECHANICS OF ENGINEERING. stone arch, and explains those already given. $AB$ is the "springing-line." 345. Mortar and Friction.—As common mortar hardens very slowly, no reliance should be placed on its tenacity as an element of stability in arches of any considerable size; though hydraulic cement, used in this place, or ordinary mortar can sometimes be depended on. Friction, however, between the surfaces of contiguous voussoirs, plays an essential part in the stability of an arch, and will therefore be considered. The stability of vousoir-arches must :—be made to depend on the resistance of the voussoirs to compression and to sliding upon each other; as also of the blocks composing the piers, the foundations of the latter being firm. 346. Point of Application of the Resistant Pressure between two consecutive voussoirs; (or pier blocks). Applying Navier's principle (as in flexure of beams) that the pressure per unit area on a joint varies uniformly from the extremity under greatest compression to the point of least compression (or of no compression); and remembering that no pressure exists (i.e., zero pressure) at points where they might in a curved beam we may represent the pressure per unit area at successive points of a joint (from the intrados towards the extrados, or vice versa) by the ordinates of a straight line, forming the surface of a trapezoid or triangle, in which figure the foot of the ordinate of the centre of gravity is the point of application of the resultant pressure. Thus, where the least compression is supposed Fig. 87. Fig. 88. Fig. 87. Fig. 88. MASONRY ARCHES. 423 to occur at the intrados $A$, Fig. 375, the pressures vary as the ordinates of a trapezoid, increasing to a maximum value at $B$, in the extrados. In Fig. 376, where the pressure is zero at $B$, and varies as the ordinates of a triangle, the resultant pressure acts within the middle third of $A$. Similarly in Fig. 377, it acts one-third the joint-length from $B$. Hence, when the pressure is not zero at either edge the resultant pressure acts within the middle third of the joint. Whereas, if the resultant pressure falls without the middle third, it shows that a portion $A_2$ of the joint, see Fig. 378, receives no pressure, i.e., the joint tends to open along $A_2$. Therefore that no joint tends to open, the resultant pressure must act within the middle third. It must be understood that the joint surfaces here dealt with are rectangles, seen edgewise in the figures. 347. Friction.—By experiment it has been found the angle of friction (see § 156) for two contiguous voussoirs of stone is about 30°. The coefficient of friction is $f = \tan 30°$. Hence if the direction of the pressure exerted upon a voussoir by its neighbor makes an angle $\alpha$ less than 30° with the normal to the joint surface, there is no danger of rupture of the arch by the sliding of one on the other. (See Fig. 379.) 348. Resistance to Crushing.—When the resultant pressure falls at its extreme allowable limit, viz.: the edge of the middle third, the pressure per unit of area at $a$, Fig. 380, is double the mean pressure per unit of area. Hence, in designing an arch of masonry, we must be assured that at every joint (taking 10 as a factor of safety) \{ Double the mean press-ure per unit of area \} must be less than $\frac{1}{10} Q$ A diagram showing a trapezoidal section of an arch with points labeled A, B, C, D. Fig. 375 A diagram showing a triangular section of an arch with points labeled A, B, C. Fig. 376 A diagram showing a rectangular section of an arch with points labeled A, B, C. Fig. 377 A diagram showing a trapezoidal section of an arch with points labeled A, B, C. Fig. 378 A diagram showing a rectangular section of an arch with points labeled A, B, C. Fig. 379 424 MECHANICS OF ENGINEERING. C being the ultimate resistance to crushing, of the material employed (§ 201) (Modulus of Crushing). Since a lamina one foot thick will always be considered in what follows, careful attention must be paid to the units employed in applying the above tests. EXAMPLE.--If a joint is 3 ft. by 1 foot, and the resultant pressure is 22.5 tons the mean pressure per sq. foot is $$p = \frac{22.5}{3 \times 1} = 7.5 \text{ tons per sq. foot}$$ ∴ its double=15 tons per sq. foot=208.5 lbs. sq. inch, which is much less than $\frac{1}{m}$ of $C$ for most building stones; see § 303, and below. At joints where the resultant pressure falls at the middle, the max. pressure per square inch would be equal to the mean pressure per square inch; but for safety it is best to assume that, at times, (from moving loads, or vibrations) it may move to the edge of the middle third, causing the max. pressure to be double the mean (per square inch). Gen. Gillmore's experiments in 1876 gave the following results, among many others:
SIZE OF BUILDING STONE. C IN LBS PER SQ. INCH.
Berea sand-stone, 2-inch cube, - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 8955
Linestone, Sebastopol, 2-inch cube (calcare), - 11720
Linestone from Cen., France, - 3650
Marble, Vermont, 2-inch cube, - 8000 to 13000
Granite, New Hampshire, 2-inch cube, - 15700 to 24000
349. The Three Conditions of Safe Equilibrium for an arch of unenamelled voussoirs. Recapitulating the results of the foregoing paragraphs, we may state, as follows, the three conditions which must be satisfied at every joint of arch-ribbed pier, for each any possible combination of loads upon the structure: (1). The resultant pressure must pass within the middle-third. (2). The resultant pressure must not make an angle > 30° with the normal to the joint. (3). The mean pressure per unit of area on the surface ARCH OF MASONRY. 425 of the joint must not exceed 1/2 of the Modulus of crush- ing of the material. 350. The True Linear-Arch, or Special Equilibrium Polygon; and the resultant pressure at any joint. Let the weight of each voussoir and its load be represented by a vertical force passing through the centre of gravity of the two, as in Fig. 381. Taking any two points, say $A$ and $B$, being in the first joint and $B$ in the last; also a third point, $p$, in the crown joint (supposing such to be there, although gener- ally a key-stone occupies the crown), through these three points can be drawn [§ 341] an equilibrium polygon for the load given; suppose this equil. polygon nowhere passes outside of the arch-ring (the arch-ring being the por- tion between the intrados $a$, $m$, and the (dotted) extrados $m'a'$ intersecting the joints at $b$, $c$, etc. Evidently if such be the case, and small metal rods (not round) were insert- ed at $A$, $b$, $c$, etc., so as to separate the arch-stones slighty, the arch would stand, though in unstable equilibrium, the piers being firm; and by a different choice of $A$, $p$, and $B$, it might be possible to draw other equilibrium poly- gons with segments cutting the joints within the arch- ring, if they were inserted; but if these segments cut the inter- sections the arch would again stand (in unstable equilibrium). In other words, if an arch stands, it may be possible to draw a great number of linear arches within the limits of the arch-ring, since three points determine an equilibrium polygon (or linear arch) for given loads. The question arises then: which linear arch is the locus of the actual re- sultant pressures of the successive joints? (Considering that we have an elastic curved beam inserted in firm piers (i.e., the blocks at the springing-line) Fig. 381. 426 MECHANICS OF ENGINEERING. are incapable of turning) and having secured a close fit at all joints before the centering is lowered, the most satisfactory answer to this question is given in Prof. Greene's "Arches," p. 131; viz., to consider the arch-ring as an arch rib of fixed ends and no hinges; see § 380 of next chapter; but the lengthy computations there employed (and the method demands a simple algebraic curve for the arch) may be most advantageously replaced by Prof. Huyghen's "Theorie der Architektonischen Statik" in Graphical Statics," published in Van Nostrand's Magazine for 1877, which applies to arch curves of any form. This method will be given in a subsequent chapter, on Arch Ribs, or Curved Beams; but for arches of masonry a much simpler procedure is sufficiently exact for practical purposes and will now be presented. If two elastic blocks of an arch-ring touch at one edge, Fig. 392, their adjacent sides making a small angle with each other, and are then gradually pressed more and more forcibly together at the edge $m$, as the arch-ring settles, the centering being gradually lowered, the surface of contact becomes larger and larger, from the compression which ensues (see Fig. 393); while the resultant pressure between the blocks, first applied at the middle of the joint, is afterwards increased under the middle of the joint in the mutual adjustment of the arch-stones. With this in view we may reasonably deduce the following theory of the location of the true linear arch (sometimes called the "line of pressures" and "curve of pressure") in an arch under given loading and with firm pier. (Whether the piers are really unyielding, under the oblique thrusts at the springing-line, is a matter for subsequent investigation.) 31. Location of the True Linear Arch—Granted that the voussoirs have been closely fitted to each other over the A diagram showing two elastic blocks of an arch-ring touching at one edge, with their adjacent sides making a small angle with each other. Fig. 392 A diagram showing two elastic blocks of an arch-ring touching at one edge, with their adjacent sides making a small angle with each other. Fig. 393 ARCH OF MASONRY. 427 centering (sheets of lead are sometimes used in the joints to make a better distribution of pressure); and that the piers are firm; and that the arch can stand at all without the centering; then we assume that in the usual accomodation of the arch, as it is constructed low-ered, the resultant of the pressures distributed over any joint, if at first near the extreme edge of the joint, advances nearer to the middle as the arch settles to its final position of equilibrium under its load; and hence the follow-ing 352. Practical Conditions. I. If for a given arch and loading, with firm piers, an equilibrium polygon can be drawn (by proper selection of the points $A$, $p$, and $B$, Fig. 381) entirely within the mid-de third of the arch ring, not only will the arch stand, but the resultant pressure at every joint will be within the middle third (Condition 1, § 349); and among all possible equilibrium polygons which can be drawn within the mid-de third of the arch-ring, the one which is nearest nearly con-sides with the middle line of the arch-ring. II. If (with firm piers, as before) no equilibrium poly-gon can be drawn within the middle third, and only one within the arch-ring at all, the arch may stand, but chipping and spalling are likely to occur at the edges of the joints. The design should - be altered. III. If no equilibrium polygon can be drawn within the arch-ring, the design of either the arch or the loading must be changed; since, although the arch may stand, from the resistance of the spaned walls, such a stability must be looked for in no other part of any construction in any large important structure. (Very frequently, in small arches of brick and stone, as they occur in buildings, the cement is so tenacious that the whole structure is virtu-ally a single continuous mass.) When the "true" linear arch has once been determined, the amount of the resultant pressure on any joint is given by the length of the proper ray in the force diagram. 428 MECHANICS OF ENGINEERING. ARRANGEMENT OF DATA FOR GRAPHIC TREATMENT. 353. Character of Load.--In most large stone arch bridges the load (permanent load) does not consist exclusively of masonry up to the road-way but partially of earth filling above the masonry, except at the faces of the arch where the span-ral walls serve as retaining walls to hold the earth. (Fig. 384). If the intrados is a half circle or half- Fig. 384. Fig. 385. ellipse, a compactly-built masonry backing is carried up beyond the springing-line to $AB$ about $60^\circ$ to $45^\circ$ from the crown, Fig. 385 ; so that the portion of arch ring below $AB$ may be considered as part of the abutment, and thus $AB$ is the virtual springing line of the arch. Sometimes to save filling, small arches are built over the haunches of the main arch, with earth placed over them, as shown in Fig. 386. In any of the preceding cases Fig. 386. Fig. 387. it is customary to consider that, on account of the bonding of the stones in the arch shell, the loading at a given distance from the crown is uniformly distributed over the width of the roadway. ARCHES OF MASONRY. 429 354. Reduced Load-Contour.—In the graphical discussion of a proposed arch we consider a lamina one foot thick, this lamina being vertical and $\frac{1}{7}$ to the axis of the arch; i.e., the lamina is on the symmetrical plane. For graphical treatment equal areas of elevation (see graph) of this lamina must represent equal weights. Taking the material of the arch-ring as a standard, we must find for each point $p$ of the extrados an imaginary height $s$ of the arch-ring material, which would give the same pressure (per running horizontal foot) at that point as that due to the actual load above that point. A number of such ordinates, each measured vertically upward from the extra- dos determine points in the "Reduced Load-Contour", i.e., the imaginary line, $AM$, the area between which and the extrados of the arch-ring represents a homogeneous load of the same density as the arch-ring, and equivalent to the actual load (above extrados), vertical by vertical. 355. Example of Reduced Load-Contour—Fig. 388. Given an arch-ring of granite (heaviness = 170 lbs per cubic foot) with a dead load of rubble (heat = 140) and earth (heat = 100), distributed as in figure. At the point $p$, of the extrados, the depth $h$ feet of rubble is equivalent to a depth of $\frac{1}{7} \times h$ feet of earth. The weight of 6 feet of granite is equivalent to $\left(\frac{1}{7} \times 6\right) \times 3.5$ feet of granite. Hence the Reduced Load-Contour has an ordinate, above $p$, of 7.6 feet. That is, for each of several points of the arch-ring extrados reduce the rubble ordinate in the ratio of $170 : 140$, and the earth ordinate in the ratio $170 : 100$ and add the re- sults, setting off the sum vertically from the points in the extrados*. In this way Fig. 389 is obtained and the area *This is most conveniently done by graphic, thus: On a right-line set off (if necessary parts of any convenient magnitude.) Call this distance $OA$. Through $O$ draw another right line parallel to $OA$. Set off $OB$ equal to $\frac{1}{14}$ for the rubble; or $\frac{1}{10}$ for the earth of the same equal part. Join $AB$. From $O$ toward $A$ set off all rubble ordinates to be reduced, each being equal to $\frac{1}{14}$ times its own length. From $O$ toward $B$ set off all earth ordinates to be reduced, each being equal to $\frac{1}{10}$ times its own length. The reduced ordinates will be the respective lengths from $O$, along $OB$ to the intersections of these parallels. With the dividers. 430 MECHANICS OF ENGINEERING. A diagram showing two arches with different load contours. Fig. 386. A diagram showing two arches with different load contours. Fig. 387. there given is to be treated as representing homogeneous granite one foot thick. This, of course, now includes the arch-ring also. $AB$ is the "reduced load-contour." 386. Live Loads.—In discussing a railroad arch bridge the "live load" (a train of locomotives, e.g., to take an extreme case) can not be disregarded, and for each of its positions we have a separate Reduced Load-Contour. EXAMPLE.—Suppose the arch of Fig. 388 to be 12 feet wide (not including spandrel walls) and that a train of locomotives weighing 5,000 lbs. per running foot of the track covers one half of the span. Uniformly * distributed laterally over the width, 12 ft., this rate of loading is equivalent to a mass of stone of one foot high and a heahtness of 250 lbs. per cubic ft., i.e., a mass of stone of 14 ft. of granite masonry (since $\frac{1}{2} \times 12 = 14$) over the half span considered. Hence from Fig. 390 we obtain Fig. 391 in an obvious manner. Fig. 391 is now ready for graphic treatment. A diagram showing two arches with different load contours. Fig. 390. A diagram showing two arches with different load contours. Fig. 391. 387. Piers and Abutments.—In a series of equal arches the pier between two consecutive arches bears simply the weight of the two adjacent semi-arches, plus the load in * It is the earth filling is shallow, the lamton directly under the track probably receive a greater pressure than the others. ARCHES OF MASONRY. 431 mediately above the pier, and . . . does not need to be as large as the abutment of the first and last arches, since these latter must be prepared to resist the oblique thrusts of their arches without help from the thrust of another on the other side. In a very long series of arches it is sometimes customary to make a few of the intermediate piers large enough to act as abutments. These are called "abutment piers," and in case one arch should fall, no others would be lost except those occurring between the same two abutment piers as the first. See Fig. 392. A is an abutment-pier. A diagram showing three arches with an abutment-pier (A) between them. Fig. 392. GRAPHICAL TREATMENT OF ARCH. 358.—Having found the "reduced load-contour," as in preceding paragraphs, for a given arch and load, we are ready to proceed with the graphic treatment, i.e., the first given, or chosen, form of the arch whose arch-ring is to be investigated with regard to stability. It will be necessary to treat, separately, a lamina under the spandrel wall, and one under the interior loading. The constructions are equally well adapted to arches of all shapes, to Gothic as well as circular and elliptical. 359.—Case I. Symmetrical Arch and Symmetrical Loading. (The "steady" (permanent) or "dead" load on an arch is usually symmetrical.) Fig. 393. From symmetry we need A diagram showing a symmetrical arch with a reduced load-contour. Fig. 393. A diagram showing a symmetrical arch with a reduced load-contour. Fig. 394. A diagram showing a symmetrical arch with a reduced load-contour. Fig. 395. 432 MECHANICS OF ENGINEERING. deal with only one half (say the left) of the arch and load. Divide this semi-arch and load into six or ten divisions by vertical lines; these divisions are considered as trapezoids except the last one, $LX$ (a convenient whole number of feet) except the last one, $LX$, next the abutment, and this is a pentagon of a different width $b_o$ (the remnant of the horizontal distance $LC$). The weight of masonry in each division is equal to (the area of division)$(\text{unity thickness of lamina}) \times (\text{weight of a cube unit of arch-ring}).$ For example for a division having an area of 20 sq. feet, and composed of masonry weighing 160 lbs. per cubic foot, we have $20 \times 1 \times 160 = 3200$ lbs., applied through the centre of gravity of the division. The area of trapezoid, Fig. 394, $\frac{1}{2}(A + B)$, and its centre of gravity may be found, Fig. 395, by the construction of Prob. 5, § 26; or by § 27a. The weight of the pentagon $LN$, Fig. 393, and its line of application (through centre of gravity) may be found by combining results for the two trapezoids into which it is divided by a vertical through point $S$, Fig. 394. Since the weights of the respective trapezoids (excepting $LN$) are proportional to their middle vertical intercepts (such as $\frac{1}{2}(h_1 + h_2)$ Fig. 394) these intercepts (transferred with the dividers) may be used directly to form the load-line, Fig. 396, or proportional parts of them if more convenient. The force scale, which this implies, is easily computed, and a proper length calculated to represent the weight of the odd division $LN$; i.e., 1...2 on the load-line. Now consider $A$, the middle point of the abutment joint, Fig. 396, as the starting point of an equilibrium polygon (or abutment of a linear arch) for a given loading, and require that this equilibrium polygon shall pass through $p$, the middle of the crown joint, and through the middle of the abutment joint on the right (not shown in figure). Proceed as in § 342, thus determining the polygon $Ap$ for the half-arch. Draw joints in the arch-ring through those points where the extrados is intersected by the ver- A diagram showing a trapezoidal section of an arch with various measurements and labels. ARCHES OF MASONRY. 433 A diagram showing the division of an arch into segments. The left side shows a section of the arch with the crown labeled "C" and the point of application of the load labeled "O". The right side shows a similar section but with the load applied at a different point, labeled "P". Fig. 367. Fig. 368. tical separating the divisions (not the gravity verticals). The points in which these joints are cut by the segments of the arch are shown in Fig. 367, are (very nearly, if the joint is not more than 60° from p, the crown) the points of application in these joints, respectively, of the resultant pressures on them, (if this is the “true linear arch” for this arch and load) while the amount and direction of each such pressure is given by the proper ray in the force-diagram. If at any joint so drawn the linear arch (or equilibrium polygon) passes outside the middle third of the arch-ring, the portion or portions of the ring may be increased (within the middle third) to find if possible a linear arch which keeps within limits at all joints. If this is found impossible, the thickness of the arch-ring may be increased at the abutment (giving a smaller increase toward the crown) and the desired result obtained ; or a change in the distribution or amount of the loading, if allowable, may gain this object. If but one linear arch can be drawn within the middle third, it may be considered the “true” one; if several, the one most nearly co-inciding with the middles of the joints (see § 301 and 302) is so considered. 300.—Case II. Unsymmetrical Loading on a Symmetrical Arch: (e.g., arch with live load covering one half-span as in Figs. 390 and 391). Here we must evidently use a full force diagram, and the full elevation of the arch-ring and load. 433 434 MECHANICS OF ENGINEERING. See Fig. 398. Select three points $A$, $p$, and $B$, as follows, to determine a trial equilibrium polygon: Select $A$ at the lower limit of the middle third of the A diagram showing a triangular section of an arch with labeled points A, p, B, and m. Fig. 398. abutment-joint at the end of the span which is the more heavily-loaded ; in the other abutment-joint take $B$ at the upper limit of the middle third; and take $p$ in the middle of the third. Draw a triangle (i.e., a linear arch) through these three points for the given set of loads, and if it does not remain within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$, within the middle third, try other positions for $A$, $p$, and $B$. Very frequently it is not necessary to draw more than one linear arch, for a given loading, for even if one could be drawn nearer the middle of the arch-ring than the first, this fact would immediately appear on mere inspection ; but those already drawn (if within middle third) will furnish values sufficiently accurate for the pressures on the respective joints, and their directions angles. 390a.—The design for the arch-ring and loading is not to be considered satisfactory until it is ascertained that for the dead load and any possible combination of live-load (in addition) the pressure at any joint is ARCHES OF MASONRY. 435 (1.) Within the middle third of that joint; (2.) At an angle of < 30° with the normal to joint-surface. (3.) Of a mean pressure per square inch not > than 1/8 of the ultimate crushing resistance. (See § 348.) § 361. Abutments.—The abutment should be compactly and solidly built, and is then treated as a single rigid mass. The pressure of the lowest voussoir upon it (considering lamina one foot thick) is given by the proper ray of the force diagram (see Fig. 399), at the point of intersection. The stability of the abutment will depend on the amount and direction of the resultant obtained by combining that pressure $P$, with the weight $G$ of the abutment and its load, see Fig. 399. Assume a probable width, $RS$ for the abutment and compute the weight $G$ of the corresponding abutment $OBRS$ and $MNBO$, and find the centre of gravity of the whole mass $C$. Apply $G$ in the vertical plane through this point, and draw a line at right angles to it at point $D$. The resultant $P$ should not cut the base $RS$ in a point beyond the middle third (or, if this rule gives too massive a pier, take such a width that the pressure per square inch at $S$ shall not exceed a safe value as computed from § 362). After one or two trials a satisfactory width can be obtained. We should also be assured that the area $PDLS$ is less than 30°. This may be tested by placing RS against it and testing as if each were, in turn, the lowest base, and if necessary may be inclined (like $m_n$) to prevent slipping. On no joint should the maximum pressure per square inch be > than 1/8 the crushing strength of the cement. Abutments of firm natural rock are of course to be preferred where they can be had. If water penetrates under an abutment its buoyant effort lessens the weight of the latter to a considerable extent. Fig. 399. We should also be assured that the area $PDLS$ is less than 30°. This may be tested by placing RS against it and testing as if each were, in turn, the lowest base, and if necessary may be inclined (like $m_n$) to prevent slipping. On no joint should the maximum pressure per square inch be > than 1/8 the crushing strength of the cement. Abutments of firm natural rock are of course to be preferred where they can be had. If water penetrates under an abutment its buoyant effort lessens the weight of the latter to a considerable extent. 436 MECHANICS OF ENGINEERING. 382. Maximum Pressure Per Unit of Area When the Resultant Pressure Falls at Any Given Distance from the Middle; according to Navier's theory of the distribution of the pressure; see § 346. Case I. Let the resultant pressure $P$, Fig. 400, (a), Fig. 400. fall within the middle third, a distance = $nd$ ($< \frac{1}{6} d$) from the middle of joint ($d$ = depth of joint). Then we have the following relations: $p$ (the mean press. per sq. in.), $p_m$ (max. press. per sq. in.), and $p_a$ (least press. per sq. in.) are proportional to the lines $h$ (middle base), $a$ (middle height), and $c$ (middle base), respectively, of a trapezoid, Fig. 400, (b), through whose centre of gravity $P$ acts. But (§ 36) $$nd = \frac{d - c}{a + c}, \text{i.e., } n = \frac{d - c}{a + c} = \frac{a - h}{h}$$ ∴ $p_m = p(6n + 1)$. Hence the following table: If $nd = \frac{1}{6} d$, then the max. press. $p_m = 2 \times \frac{1}{6} d$, times the mean pressure. Case II. Let $P$ fall outside the mid. third, a distance = $nd$ ($> \frac{1}{6} d$) from the middle of joint. Here, since the joint is not considered capable of withstanding tension, we have a triangle, instead of a trapezoid. Fig. 401. First compute the mean press. per sq. in. $$p = \frac{P (\text{lbs.)}}{(1 - 2n) 18 d \text{ inches}}$$ or from this table: (lamina one foot thick). Fig. 401. ARCHES OF MASONRY. 437
For nd = A d A d A d A d A d A d
p = $\frac{1}{10}$ P $\frac{1}{8}$ P $\frac{1}{6}$ P $\frac{1}{4}$ P $\frac{1}{2}$ P infinity.
(at in inches and P is lb., with each lamina 1 ft. in thickness.) Then the maximum pressure (at A, Fig. 401) $p_{max} = 2p$, becomes known, in lbs. per sq. in. 362a. Arch-ring under Non-vertical Forces.—An example of this occurs when a vertical arch-ring is to support the pressure of a liquid on its extrados. Since water-presures are always at right angles to the surface pressed on, these pressures on the extradosal surface of the arch-ring form a system of non-parallel forces which are normal to the curve of the extrados at their respective points of application, and lie in parallel vertical planes passing through the lamina. We now assume that the extradosal surface is a cylinder (in the most general sense) whose rectilinear elements are $\gamma$ to the faces of the lamina. If, then, we divide the length of the extrados, from crown to each abutment, into six to ten parts, the respective pressures on the corresponding surfaces are obtained by multiplying the area of each by the depth of its centre of gravity from the upper free surface of the liquid, and this product by the weight of a unit of volume of the liquid; and each such resultant is then added together to give the weight of the centre of the area. Finally, if we find the resultant of each of these pressures and the weight of the corresponding portion of the arching, these resultants form a series of non-vertical forces in a plane, for which an equilibrium polygon can then be passed through three assumed points by § 378a, these three points being taken in the crown-joint and the two abutment-joints. As to the “true linear arch” see § 559. As an extreme theoretical limit it is worth noting that if the extradosal surface of the arch is a concentric circle; if the weights of the voussoirs are neglected ; and if the rise of the arch is very small compared with the depth of the crown below the water surface, then the circularcentre-line of the arch-ring is the “true linear arch.” A diagram showing an arch ring with extradosal surface divided into segments and pressures applied at various points along its length. 438 MECHANICS OF ENGINEERING. CHAPTER XL ARCH-RIBS. 364. Definitions and Assumptions.—An arch-rib (or elastic-arch), as distinguished from a block-work arch) is a rigid curved beam, either solid, or built up of pieces like a truss (and then called a braced arch) the stresses in which, under a given loading and with prescribed mode of support is here proposed to determine. The rib is supposed symmetrical about its axis or middle line, and the Moment of Inertia of any cross section is understood to be referred to a gravity axis of the section, which (the axis) is perpendicular to the said vertical plane. It is assumed that in its strained condition under a load, the shape of the rib differs so little from its form when unstrained that the change in the abscissa or ordinate of any point in the rib axis (a curve) may be neglected when added (algebraically) to the co-ordinates of a point, and that the radius of a truss-section are small compared with the radius of curvature at any part of the curved axis, and with the span. 365. Mode of Support.—Either extremity of the rib may be hinged to its pier (which gives freedom to the end-tangent-line to turn in the vertical plane of the rib when a load is applied), or may be fixed, i.e., so built-in, or bolted rigidly to the pier, that the end-tangent-line is incapable of changing its position when a load is applied. A hinge may be inserted anywhere along the rib, and of course ARCHI RIBS. 439 destroys the rigidity, or resistance to bending at that point (A hinge having its pin horizontal $T$ to the axis of the rib is meant). Evidently no more than three such hinges could be introduced along an arch-rib between two piers; unless it is to be a hanging structure, acting as a suspension-cable. 366. Arch Rib as a Free Body.—In considering the whole rib free it is convenient, for graphical treatment, that no section be conceived made at its extremities, if fixed ; hence in dealing with that mode of support the end of the rib will be considered as having a rigid prolongation reaching to a point vertically above or below the pier junction, an unknown distance from it, and there acted on by a force of such unknown amount and direction as to preserve the actual extension of the rib and its tangent line in the same position and direction as they really are. As an illustration of this Fig. 402 shows free an arch rib. ONB, with its extremities $O$ and $B$, fixed in the piers, with no hinges, and bearing two loads $P_1$ and $P_2$. The other case of the system holding it in equilibrium are the horizontal and vertical components, of the pier reactions ($H_1$, $V_1$, $H_2$, and $V_2$), and in this case of fixed ends each of these two reactions is a single force not intersecting the end of the rib, but cutting the vertical through the end in some point $F$ (on the left; and in $G$ on the right) at some vertical distance $c$, (or $\delta$), from the end. Hence the utility of these imaginary prolongations $OF$, and $BG$, when their being supposed removed. Compare Figs. 348 and 350. The imaginary points, or hinges, $F$ and $G$, will be called abutments being such for the special equilibrium polygon A diagram showing an arch rib with its extremities O and B fixed in piers, with no hinges, and bearing two loads P1 and P2. The other case of the system holding it in equilibrium are the horizontal and vertical components of the pier reactions H1, V1, H2, and V2. Fig. 402 440 MECHANICS OF ENGINEERING. (dotted line), while $O$ and $B$ are the real ends of the curved beam, or rib. In this system of forces there are five unknowns, viz.: $V$, $V_{a}$, $H = H_{a}$, and the distances $c$ and $d$. Their determination by analysis, even if the rib is circular, are in extremely difficult cases, to be solved by graphical solutions (Prof. Eddy's method; see § 350 for reference), it is comparatively simple and direct and applies to any shape of rib, and is sufficiently accurate for practical purposes. This method consists of constructions leading to the location of the "special equilibrium polygon" and its force diagram. In case the rib is hinged to the piers, the reactions of the latter act through these hinges, Fig. 403, i.e., the abscissas of the special equilibrium polygon represent the shears of the rib $O$ and $B$, and for a given rib and load the unknown quantities are only three $V$, $V_{a}$, and $H$ (strictly there are four; but $\Delta X = 0$ gives $I_{c} = H$). The solution by analytics is possible only for ribs of simple algebraic curves and is long and cumbersome; whereas Prof. Eddy's graphic method is comparatively brief and simple and is applicable to any shape of rib whatever. 367. Utility of the Special Equilibrium Polygon and its force diagram. The use of locating these will now be illustrated (See § 332). As proved in §§ 332 and 334 the compression in each "rod" or segment of the "special equilibrium polygon" in the anti-stress resultant of the cross sections in the corresponding portion of the beam, rib, or other structure, the value of this compression (in lbs. or tons) being measured by the length of the corresponding curve diagram. Suppose that these way (to be explained subsequently) the special equilibrium polygon and its force diagram have been drawn for the arch-rib in Fig. 404 having fixed ends $O$ and $B$, and no hinges ; required the elastic stresses in any cross-section of the rib as at m. Let the Fig. 403. 403 ARCH BIBS. 443 Fig. 404. scale of the force-diagram on the right be 200 lbs. to the inch, say, and that of the space-diagram (on the left) 30 ft. to the inch. The cross section $m$ lies in a portion $TK$, of the rib, corresponding to the rod or segment $bc$ of the equilibrium polygon; hence its anti-stress-reaction is a force $R_2$ acting in the line $bc$, and of an amount given in the force-diagram. Now $R_2$ is a resultant of $T_2$ and $J_2$, which are the elastic forces at $m$ forming a system in equilibrium, shown in Fig. 405; the portion $FO/Tm$ being considered free. Hence Fig. 405. taking the tangent line and the normal at $m$ as axes we should have $\Sigma(\tan\text{ compa})=0$; $\Sigma(\text{ norm. compa})=0$; and $\Sigma(\text{ moma. about gravity axis of the section at } m)=0$, and could thus find the unknowns $p_1$, $p_2$, and $J$, which appear in the expressions $p_1F$ the thrust, $\frac{p_1}{c}$ the moment of Fig. 406. 442 MECHANICS OF ENGINEERING. the stress-couple, and $J$ the shear. These elastic stresses are classified as in § 295, which see. $p_{1}$ and $p_{2}$ are the per square inch, $J$ is $Be$, $e$ is the distance from the horizontal gravity axis of the section to the outermost element of area, (where the compression or tension is $p_1$ lbs. per sq. in., as due to the stress-couple alone) while $y$ is the "movement" of this point from the neutral gravity axis (See §§ 247 and 295; also § 85). Graphics, however, give us a more direct method, as follows: Since $R_m$ in the line bc, is the equivalent of $V$, $H$, and $P$, the stresses at m will be just the same as if $R_m$ acted directly upon a lateral prolongation of the rib at T (to intersect & Fig. 405) as shown in Fig. 406, this prolongation $Tb$ taking the place of $TOP$ in Fig. 405. The force diagram is also reproduced here. Let $\alpha$ denote the length of the $\gamma$ from m's gravity axis upon oc; and s the vertical intercept between m and bc. For this imaginary free body, we have, $$\text{from } \gamma (\tan c \cos m)=0, R_s \sin e = J$$ and from $\gamma$ (norm. compn.)=0, $R_s \sin e = J$ while from $\gamma$ (mom. about) $\frac{1}{e}$ we have $R_{es} = p_1 f$ the gravity axis of m)$=0.$ But from the two similar triangles (shaded; one of them is in force diagram) $\alpha : z :: H; R_s \cdot R_{es} = Hs$, whence we may rewrite these relations as follows (with a general statement), viz: If the Special Equilibrium Polygon and Its Force Diagram Have Been Drawn for a given arch-rib, of given mode of support, and under a given loading, then in any cross-section of the rib, we have ($F^{\prime}$ = area of fig.) (L) The Thrust, i.e., $p_1 f =$ The projection of the proper ray (of the force diagram) up on the tangent line of the rib drawn at the given section. 9/22 ARCH RIBS. -43- (2.) The Shear, i.e., $J = \frac{1}{e}$ (upon which depends the shearing stress in the web). (See §§ 253 and 256). The projection of the proper ray (of the force diagram) upon the normal to the rib curve at the given section. (3.) The Moment of the stress couple, i.e., $\frac{F_1}{e} = H$. The product ($H e$) of the $H$ (or pole-distance) of the force-diagram by the vertical distance of the ground axis of the section from its special equilibrium polygon. By the "proper ray" is meant that ray which is parallel to the segment (of the equil. polygon) immediately under or above which the given section is situated. Thus in Fig. 404, the proper ray for any section on $TK$ is $R_1$; on $KH$, $R_2$; on $TO$, $R_3$. The projection of a ray upon any given tangent or normal, is easily found by drawing through each end of the ray a line perpendicular to the tangent (or normal); the length of this line is $TQ'$ on the tangent, and $TQ$ is the force required (by the scale of the force diagram). We may thus construct a shear diagram, and a thrust diagram for a given case, while the successive vertical intercepts between the rib and special equilibrium polygon form a moment diagram. For example of this of a point $m$ $\frac{1}{4}$ inch in a space diagram drawn to a scale of 20 feet to the inch, while $H$ measures 2.1 inches in a force diagram constructed on a scale of ten tons to the inch, we have for the moment due to shear at this point, $M_{\text{shear}} = (2.1 \times 10)$ tons x $\left[\frac{1}{4} \times 20\right]$ ft. = 210 ft. tons. 388.—It is thus seen how a location of the special equilibrium polygon, and the lines of the corresponding force-diagram, lead directly to a knowledge of the stresses in all the cross-sections of the curved beam under consideration, bearing a given load; or, vice versa, leads to a statement of conditions to be satisfied by the dimensions of the rib, for proper security. It is here supposed that the rib has sufficient lateral 444 MECHANICS OF ENGINEERING. bracing (with others which lie parallel with it) to prevent buckling sideways in any part like a long column. Before proceeding to the complete graphical analysis of the different cases of arches, it is better to devote the next few paragraphs to developing a few analytical relations in the theory of flexure of a curved beam, and to giving some processes in "graphical arithmetic." 369. Change in the Angle Between Two Consecutive Rib Tangents when the rib is loaded, as compared with its value before loading. Consider any small portion (of an arch rib) included between two consecutive cross-sections; Fig. 407. $KHGW$ is its unstrained form. Let $EA = ds$, be the original length of this portion of the rib axis. The length of all the fibres ($i$ to rib-axis) was originally $= de$ (nearly) and the two consecutive tangent-lines, at $E$ and $A$, made an angle $= dt$ originally, with each other. While under strain, however, all the fibres are shortened equally an amount $de$, by the uniformly distributed tangential thrust, acting on them, and also by the centrifugal force acting as they are on one side or the other of the gravity axis $E$, or $J$, of the section) by the system of forces making what we call the "stress couple," among which the stress at the distance $r$ from the gravity axis $A$ of the section is called $p_r$ per square inch; so that the tangent line at $A'$ now takes the direction $AD \rightarrow A'G'$ instead of $AQ$ (we suppose the section at $E$ to remain fixed, for conveni- A diagram showing a curved beam with a rib structure. The diagram includes various labeled points and lines representing different sections of the beam. Fig. 407 $$\frac{dF}{ds} = \frac{dp_r}{dr}$$ ARCH RIBS. 445 ence, since the change of angle between the two tangents depends on the stresses acting, and not on the new position in space, of this part of the rib, and hence the angle between the tangent-line $A$ and $A'$ (original line) is moved by the couple $CA'D' \times G'H'$ to $G'H''$ is the new position of $GH$. We obtain the value of $d\phi$ as follows: That part ($d\phi$) of the shortening of the fibre at $G$, at distance $e$ from $A$ due to the force $p_dF$, is § 201 eq. (1), $d\phi = \frac{p_d e}{E}$. But, geometrically, $d\phi$ also = $d\phi_e$ $\therefore \quad d\phi_e = p_d e \quad \ldots \quad \ldots \quad \ldots \quad \ldots \quad (L)$ But, letting $M$ denote the moment of the stress-couple at section $A$ ($M$ depends on the loading, mode of support, etc., in any particular case) we know from § 295 eq. (6) that $M = \frac{P_d e}{E} \quad \ldots \quad \ldots \quad \ldots \quad (2)$ [If the arch-rib in question has less than three hinges, the equal shortening of the fibres due to the thrust (of the block in last figure) $p_dF$, will have an indirect effect on the angle $d\phi$. This will be considered later.] 370. Total Change [i.e. $\int d\phi_e$] in the Angle Between the End Tangents of a Rib, before and after loading. Take the example in Fig. 408 of a rib fixed at one end and hinged at A diagram showing a rib with a hinge at one end and another hinge at the other end. The diagram includes labels for angles $\theta_1$, $\theta_2$, and $\theta_3$, as well as points A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. Fig. 408 446 MECHANICS OF ENGINEERING. the other. When the rib is unstrained (as it is supposed to be, on the left, its own weight being neglected; it is not supposed sprung into place, but is entirely without strain) then the angle between the end-tangents has some value $\theta = \int_0^a d\theta$ the sum of the successive small angles $d\theta$ for each element $ds$ of the rib curve (or axis). After loading, [on the right, Fig. 408], this angle has increased having now a value $\theta' + \int_0^a ds = \theta + \int_0^a \frac{Mds}{EI} \quad \ldots \quad (L)$ There must be no hinge between $O$ and $B$. § 371. Example of Equation (L) in Analysis.--A straight, homogeneous, prismatic beam, Fig. 409, its own weight neglected, is fixed obliquely in a wall. After placing a load $P$ on the free end, required the angle between the end-tangents. This was zero before loading ; its value after loading is $=0 + \varphi' = 0 + \frac{1}{EI} \int_0^a Mds$ By considering free a portion between $O$ and any $ds$ of the beam, we find that $M-Px$=mom. of the stress couple. The flexure is so slight that the angle between any $dx$ and its $ds$ is still practically zero ($\S$ 364), and, $\therefore ds=dx$ sec $\alpha$. Hence, by substitution in eq. (L) we have $\varphi' = \frac{1}{EI} \int_0^a Mds = \frac{Psec\alpha}{EI} \int_0^a dx = \frac{Psec\alpha}{EI} \frac{\pi}{2}$; $\therefore \varphi' = P(\cos\alpha)^2 \frac{\pi}{2EI}$ (Compare with § 237). Fig. 409. ARCH RIBS. 447 It is now apparent that if both ends of an arch rib are fixed, when unstrained, and the rib be then loaded (within elastic limit, and deformation slight) we must have $$\int_0^a(Mdx+EI)=0,$$ since $q'=0.$ 372. Projections of the Displacement of any Point of a Loaded Rib Relatively to Another Point and the Tangent Line at the Latter.--(There must be no hinges between $O$ and $B$). Let $O$ be the point whose displacement is considered and $B$ the other point. Fig. 410. If $B$'s tangent-line is fixed while the extremity $O$ is not supported in any way (Fig. 410) then a load $P$ put on, $O$ is displaced to a new position $O_x$. Fig. 410. Fig. 411. Fig. 412. With $O$ as an origin and $OB$ as the axis of $X$, the projection of the displacement $OO_x$ upon $X$, will be called $Jx$, that upon $Y$, $Jy$. In the case in Fig. 410, $O$x's displacement with respect to $B$ and its tangent-line $BT$, is also its absolute displacement in space, since neither $B$ nor $BT$ has moved as the rib changes form under the load. In Fig. 411, however, the extremities $O$ and $B$ are both hinged to piers, or supports, the dotted line showing its form when deformed under a load. The hinges are supposed immovable, the rib being free to turn about them. Hence the absolute displacement of the changed form under a load, and the absolute displacement of $O$ is zero; but not so its displacement relatively to $B$ and its tangent $BT$, for $BT$ has moved to a new position $BT'$. To find this relative displacement conceive the new curve of the rib superposed on the old in a way that $B$ and $BT'$ may coincide with their original po- 448 MECHANICS OF ENGINEERING. sitions, Fig. 412. It is now seen that $O'$'s displacement relatively to $B$ and $BT$ is not zero but = $OD_{a}$, and has a small $\Delta x$ but a comparatively large $\Delta y$. In fact for this case of hinged ends, per se immovable, rib continuous between them, and deformation slight we shall write $\Delta z$ = zero as compared with $\Delta y$, the case $X$ passing through $OB$). 373. Values of the X and Y Projections of $O'$'s Displacement Relatively to $B$ and its Tangent; the origin being taken at $O$. Fig. 413. Let the co-ordinates of different points $E$, $D$, $C$, etc., of the rib, referred to $O$ and an arbitrary $X$ axis, be $x$ and $y$, their radial distances from $O$ being $u$ (i.e., $u$ for $C$, $w$ for $D$, etc.; in general, $u = OED$ is the mean strain form of the rib (e.g., the form it would assume if it lay flat on its side on a level platform, under no straining forces), while $OE', D'C'B$ is its form under some loading, i.e., under strain. (The superposition above mentioned (§ 372) is supposed already made if necessary, so that $BT$ is tangent at $B$ to both forms). Now conceive the rib $OB$ to pass into its strained condition by the successive bending of each de in turn. The straining or bending of the first de, $BC$, through the small angle $\delta e_1$ (dependent on the moment of the stress couple at $C$ in the strained condition) causes the point finite distance OC to turn about $C$ as a centre through the same small angle $\delta e_1$; hence the point $O$ describes a small linear arc OO' = $\delta e_1 r_0$, whose radius is $u$, the hypotenuse of the $x$ and $y$ of $C$, and whose value . . . is $\delta e_1 u = \delta e_1 w$. Next let the section $D$, now at $D'$, turn through its proper angle $\delta e'$ (dependent on its stress-couple) carrying ARCHI RIBS. 449 with it the portion $D'O'$, into the position $D'O''$, making $O'$ describe a linear arc $O'O''=(bo')'=w'do'$, in which $w$=the hypotenuse on the $x'$ and $y'$ (of $D$), (the deformation is so slight that the co-ordinates of the different points referred to $O$ and $X$ are not appreciably affected). Thus, each section having been allowed to turn through the angle proper to it, $O$ finally reaches its position, $O_0$, of displacement. Each successive $\delta o$, or linear are described by $O$, has a shorter radius. Let $\delta x$, ($\delta x'$), etc., represent the projections of the successive ($\delta o$) upon the axis $X$; and similarly $\delta y$, ($\delta y'$) etc., upon the axis $Y$. Then the total $X$ projection of the curved line $O \ldots O_0$ will be $$dx = \int \delta x \text{ and similarly } dy = \int \delta y \quad \ldots \quad (1)$$ But $\delta v = u\,\delta y$, and from similar right-triangles, $$\delta z : \delta v :: y : u \text{ and } dy : \delta y :: z : u \Rightarrow \delta x = y\,\delta y \text{ and } \delta y = zd\varphi;$$ whence, (see (1) and (2) of §369) $$dx = \int \delta x = \int yd\varphi = \int_{a}^{b} Myde \quad \ldots \quad (II)$$ and $$dy = \int \delta y = zd\varphi = \int_{a}^{b} Myde \quad \ldots \quad (III)$$ If the rib is homogeneous $E$ is constant, and if it is of constant cross-section, all sections being similarly cut by the vertical plane of the rib's axis (i.e., if it is a "curved prism") $I$, the moment of inertia is also constant. 374. Recapitulation of Analytical Relations, for reference. (Not applicable if there is a hinge between $O$ and $B$.) Total Change in Angle between tangent-lines $O$ and $B$ $$=\int_{a}^{b} Myde \quad \ldots \quad (L)$$ The X-Projection of $O'O''$ Displacement-line; Relative to B (the origin being at $O$) $$=\int_{a}^{b} Myde \quad \ldots \quad (IL)$$ 450 MECHANICS OF ENGINEERING. The Y-Projection of 0's Displacement, etc., as above. $$\left\{\begin{array}{l}{}\\ {}\\ \end{array}\right. = \int_0^a \frac{Mdx}{EI}$$ (III). Here $x$ and $y$ are the co-ordinates of points in the rib-curve, $de$ an element of that curve, $M$ the moment of the stress-couple in the corresponding section as induced by the loading, or constraint, of the rib. (The results already derived for deflections, slopes, etc., for straight beams, could also be obtained from these formulae, I., II. and III. In these formulae also it must be remembered that no account has been taken of the shortening of the rib-axis by the thrust, nor of the effect of a change of temperature.) 374a. Remarks on the Properties of Equilibrium Polygons and their Force Diagrams, for Systems of Vertical Loads.—See § 335 to 343. Given a system of loads or vertical forces, $F_1$, $P_n$, etc., Fig. 414, and two abutment verti- cals, $F'$ and $G'$; if we lay off, vertically, to form a load- line, $F_1$, $P_1$, $2...$ $= F_2$, etc., select any Pole, $O_1$, and join $O_1$ $\ldots 1$, $O_1$, $2$, etc.; also, beginning at any point $F'_1$ in the vertical $F'$, if we draw $F'_1\ldots a\parallel O_1$ to intersect the line of $F'_1$; then $ab\parallel O_1$, $2$ and so on until finally a point $O_2$, in $G'$, is determined; then the figure $F'_1abcG'_1$ is an equilibrium polygon for the given loads and load verti- cals, and (§ 335) a force diagram. This former is so called because the short segments $Fa$, $ob$, etc., if considered to be rigid and imponderable rods, in a vertical plane, hinged to each other and the terminal ones to abut- ments $F'_1$ and $G'_1$, would be in equilibrium under the given loads hung at the joints. An infinite number of equilibri- Fig. 414 ARCH-RIBS. 451 rium polygons may be drawn for the given loads and abutment-vertices, by choosing different poles in the force diagram. (One other is shown in the figure; $O_0$ is its pole. ($F_1$, $G_1$ and $P_1$, $G_1$ are abutment forces.) For all of these the following statements are true: (1.) A line through the pole, I to the abutment-line cuts the load-line in the same point $n'$, whichever equilibrium polygon be used ( : any one will serve to determine $n'$.) (2.) If a vertical $CD$ be drawn, giving an intercept $x'$ in each of the equilibrium polygons, the product $H'x'$ is the same for all the equilibrium polygons. That is, (see Fig. 414) for any two of the polygons we have $$H_1 \cdot H_2 \cdot x'_1 \cdot x'_2 = H_3 \cdot H_4 \cdot x'_3 \cdot x'_4$$ (3.) The compression in each segment is given by that "ray" (in the force diagram) to which it is parallel. (a.) The pole distances $H_1$ or $H_2$ let fall from the pole upon the load-line divide it into two parts which are the vertical components of the compressions in the abutment-rods respectively (the other component being horizontal); $H$ is the horizontal component of each (and, in fact, of each of the compressions in all the other rods). The compressions in the extreme rods may also be called the abutment reactions (oblique) and are given by the extreme rays. (b.) Three Points [not all in the same segment (or rod)] determine an equilibrium polygon for given loads. Having given, then, three points, we may draw the equilibrium polygon by §341. 375. Summation of Products. Before proceeding to treat graphically any case of arch-ribs, a few processes in graphical arithmetic as it may be called, must be presented, and thus established for future use. To make summation of products of two factors in each by means of an equilibrium polygon. 452 MECHANICS OF ENGINEERING. Construction. Suppose it required to make the summa- tion $X(x)$ i. e., to sum the series $$x_1 x_2 + x_3 x_4 + \ldots$$ by graphics. Having first arranged the terms in the order of magni- tude of the $x_i$, we proceed as follows: Supposing, for illustration, that two of the $x_i$ ($a_0$ and $z_1$) are negative (dotted lines), and that all other quantities $x$ and $w$ may be of any nature whatever, anything capable of being represented by a length, laid off to scale. Fig. 416. First, in Fig. 416, lay off the $a_i$'s in their order, end to end, on a ver- tical load-line taking care to lay off $a_0$ and $a_1$ first, in their turn. Take any con- venient pole $O$; draw the rays $O\ldots 1$, $O\ldots 2$, etc.; then, having pre- viously drawn vertical lines whose horizontal distances from an extreme left-hand vertical $F^*$ are made = $x_1$, $x_2$, $x_3$, etc., respectively, we begin at any point $F$, in the verti- cal $F^*$, and draw a line (to $G^*$) to intersect the ver- tical $F^*$, and so on until the last $V'$ to intersect $F^*$, and so on fol- lowing carefully the proper order. Produce the last seg- ment ($G'\ldots G$ in this case) to intersect the vertical $F^*$ in some point $K$. Let $KF = k$ (measured on the same scale as the $x_i$'s), then the summation required is $$S^*_{(x)} = Hk.$$ $H$ is measured on the scale of the $a_i$, which need not be the same as that of the $x_i$'s; in fact the $z_i$ may not be the same kind of quantity as the $x_i$'s. [Proof.] From similar triangles $H::z_1::k_0$, $\therefore z_1x_1 = Hk_0$; and " " " " " " $H::z_2::k_0$, $\therefore z_2x_2 = Hk_0$. Fig. 417. ARCH-RIBS. 453 and so on. But $H(k_1+k_2+\text{etc.}) = H \times FK = Hk_1$. **376. Gravity Vertical.** From the same construction in Fig. 415 we can determine the line of action (or gravity vertical) of the resultant of the parallel vertical forces $z$, $e$, etc. (or loads); by prolonging the first and last segments to their intersection at $F$. The resultant of the system of vertical forces acts through $C$ and is vertical in this case; its value being = $\frac{1}{7}$ ($z$), that is, it = the length 1...7 in the force diagram, interpreted by the proper scale. It is now supposed that the $z$'s represent the areas of small finite portions of a large plane figure, we may find a gravity-line (through $C$) of that figure by the algebraic sum of all the $z$'s being applied through the centre of gravity of its own portion. Calling the distance between the verticals through $C$ and $F$, we have also $\frac{1}{7} (z) = \frac{1}{7} (z)$ because $z$ is the resultant of the $1\ldots7$. This is evident from the proportion (similar triangles) $$H : (1\ldots7) = z : k$$ 454 MECHANICS OF ENGINEERING. 376 a. Moment of Inertia (of Plane Figure) by Graphs - Fig. 416 a. $I_x = ?$ First, for the portion on right. Divide OR into equal parts each $= dx$. Let $x$, $z$, etc., be the middle ordinates of the strips thus obtained, and $x_0$ etc. their abscissas (of middle points). Then we have approximately $$I_x \text{ for } OR = dx.z_0^2 + dx.z_1^2 + \ldots \ldots = dx[(z_0)^2 + (z_1)^2 + \ldots] \text{ . . .} (1)$$ But by §375 we may construct the products $z.z_0.z_1$, etc., taking a convenient $H'$ (see Fig. 416, b), and obtain $k_z$, $k_x$ etc., such that $z.x = H'k_z$, $z.y = H'k_x$, etc. Hence eq.(1) becomes: $$I_x \text{ for } OR \approx H'(k_z.k_x + k_z.k_y + \ldots) \text{ . . .} (2)$$ By a second use of § 375 (see Fig. 416 c) we construct $l$, such that $k_z.l + k_x.l + \ldots = H''(l'')$ ($H''$ taken at convenience). . . . from eq. (2) we have finally, (approx.) $$I_x \text{ for } OR = H'H''l.zl \text{ . . .} (3)$$ For example if $OR = 4$ in., with four strips, $dx$ would = 1 in., and if $H' = 2$ in., $H'' = 2$ in., and $l = 3.2$ in., then $$I_x \text{ for } OR = 2\times2\times5.2\times1.0 = 20.8\text{ biquad inches}.$$ The $I_x$ for $OL$, on the left of $N$, is found in a similar manner and added to $I_x$ for $OR$ to obtain the total $I_x$. The position of a gravity axis is easily found by cutting the shape out of sheet metal and balancing on a knife edge; or may be obtained graphically by § 386, or 376. 377. Construction for locating a line em (Fig. 417) at $(a)$, is the polygon FG in such a position as to satisfy the two following conditions with reference to the vertical intercepts at 1, 2, 3, 4, and 5, between it and the given points 1, 2, 3, etc., of the perimeter of the polygon. ARCH-RIBS. 455 Condition I. (Calling these intercepts $x_1$, $x_2$, etc., and their horizontal distances from a given vertical $F$, $x_3$, $x_4$, etc.) $S(x)$ is to = 0; i.e., the sum of the positive $w$'s must be numerically that of the negative (which here are at 1 and 5). An infinite number of positions of $w$ will satisfy condition I. Condition II. $S(wx)$ is to = 0; i.e., the sum of the moments of the positive $w$'s about $F$ must = that of the negative $w$'s, i.e., the moment of the resultant of the positive $w$'s must = that of the resultant of the negative; and, being already shown that these two resultants must be directly opposed and equal. But the ordinates $u$ in ($a$) are individually equal to the difference of the full and dotted ordinates in ($b$) with the same $x$'s : - the conditions may be written : I. $\Sigma$ (full ord. in ($b$)) = $\Sigma$ (dotted ord. in ($b$)) II. $\Sigma$ (each full ord. in ($b$) × its $x$) = $\Sigma$ (each dotted ord. in ($b$) × its $x$), i.e., the centres of gravity of the full and of the dotted in ($b$) must lie in the same vertical. Again, by joining $vG$, we may divide the dotted ordinates of ($b$) into two sets which are dotted, and broken, respectively, in ($c$). Then, finally, drawing in ($d$), $$R,$$ the resultant of full ord. of ($c$) $$T,$$ " " " broken " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " 456 MECHANICS OF ENGINEERING. we are prepared to state in still another and final form the conditions which we must fulfil, viz. : (L) $T + T' = R$ ; and (II.) The resultant of $T$ and $T'$ must act in the same vertical as $R$. In short, the quantities $T$, $T'$, and $R$ must form a balanced system, considered as forces. All of which amounts practically to this: that if the verticals in which $T$ and $T'$ are set known and $R$ be conceived as a load supported by a horizontal beam (Fig. 417), then (as in Fig. 418) resting on piers in those verticals, then $T$ and $T'$ are the respective reactions of those piers. It will now be shown that the verticals of $T$ and $T'$ are easily found, being independent of the position of $v$: and that both the vertical and the magnitude of $R$, being likewise independent of $v$, are determined with facility in advance. For, if $b$ be shifted up or down, all the broken ordinates in $(c)$ or $(d)$ will change in the same proportion (viz. as $vF$ changes), while the dotted ordinates though altered, remain unchanged, do not change value because the shifting of piers affects neither the vertical nor the value of $T'$ nor the vertical of $T$. The value of $T$, however, is proportional to $vF$. Similarly, if $m$ be shifted, up or down, $T'$ will vary proportionally to $mF$, but its vertical, or line of action, remains the same. $T$ is unaffected in any way by the shifting of $m$. $R$, depending for its value and position on the full ordinates of $(c)$ Fig. 417, is independent of the location of $v$. We may -, proceed as follows: Let us represent $R$ graphically, in amount and position, by means of § 376. 2ndly. Determine the verticals of $T$ and $T'$ by any trial position of $v$ (call it $v_{0}m_0$), and the corresponding trial values of $T$ and $T'$ (call them $T_1$ and $T'_1$). 3rdly. By the fiction of the horizontal beam, construct ($§ 329$) or compute the true values of $T$ and $T'$, and then determine the true distances $vF$ and $mG$ by the proportions $$vF : v_0F :: T : T_1 \text{and } mG : m_0G :: T' : T'_1$$ ARCH-RIBS. 457 Example of this Fig. 418. (See Fig. 417 for $s$ and $t_c$.) From $A$ toward $B$ in (e) Fig. 418, lay off the lengths (or lines proportional to them) of the full ordinates 1, 2, etc., of ($f$). Take any pole $O_a$, and draw the equilibrium polygon ($f$) and prolong its segments to find $C$ and thus determine $R'$s vertical. $R$ is represented by $AB$. In $(g)$ [same as ($f$) but shifted to avoid complexity of lines] draw a trial $v_m$, and join it with $v_n$. Determine the sum $T_s$ of the broken ordinates (between $v_0G_1$ and $F_3G_3$) and its vertical line of application, precisely as in dealing with $R$; also $T_s$, that of the dotted ordinates (five) and its vertical. Now the true $T = R + (s + t)$ and the true $T' = R + (s + t - c)$. Hence compute $\phi = \frac{T - T'}{T} v_F$, and $\mu = \frac{T + T'}{T} m_g G_n$ and by laying off vertically upward from $F$ and $G$ respectively we determine the true values of these quantities under the conditions imposed at the beginning of this article, relating to the vertical ordinates intercepted between $m_n$ and given points on the perimeter of a polygon or curve. Note (a): If the verticals in which the intercepts lie are equidistant and quite numerous, then the lines of action of $T_s$ and $T'_s$ will divide the horizontal distance between $F$ and $G$ into three equal parts. This will be exactly true in the application of this construction to § 390. Note (b): If the verticals are not uniformly placed about a vertical line, (as will usually be the case) $v_m$, is Fig. 418 458 MECHANICS OF ENGINEERING. best drawn parallel to $FG$, for then $T_1$ and $T'$ will be equal and equidistant from said vertical line. 378. Classification of Arch-Rib, or Elastic Arches, according to continuity and modes of support. In the accompanying figures the full curves show the *unstrained* form of the rib (before any load, even its own weight, is permitted to come upon it); the dotted curve shows its shape (much exaggerated) under a given loading. Three Conditions must be given to determine the special equilibrium polygon ($\S$ 366 and 367). Class A.—Continuous rib, free to slip laterally on the piers, which have smooth horizontal surfaces, Fig. 490. This is chiefly of theoretic interest, its consideration being therefore omitted. The pier reactions are necessarily vertical, just as if it were a straight horizontal beam. Class B. Rib of Three Hinges, two at the piers and one intermediate (usually at the crown) Fig. 491. Fig. 36 also is an example of this. That is, the rib is discontinuous and of two segments. Since at each hinge the moment of the stress couple must be zero, the special equilibrium polygon must pass through the hinges. Hence as three points fully determine an equilibrium polygon for given load, the special equilibrium is drawn by § 341. Fig. 490 Fig. 491 [$\S$ 378 will contain a construction for arch-rib of three hinges, when the forces are not all vertical.] Class C. Rib of Two Hinges, these being at the piers, the rib continuous between. The piers are considered immovable, i.e., the span cannot change as a consequence of loading. It is also considered that the rib is fitted to its ARCH RIBS. 459 hinges at a definite temperature, and is then under no constraint from the piers (as if it lay flat on the ground), not even its own weight being permitted to act when it is finally put into position. When the "false works" or temporary supports are removed, stresses are induced in the rib both by its loading, including its own weight, and by a change of temperature. Stresses due to temperature may be ascertained separately and then combined with those due to loading. (Chapters A and B are not subject to temperature stresses.) Fig. 422 shows a rib of two hinges, at ends. Conceive the dotted curve (form and position under strain) to be superposed on the continuous curve (form before strain) in such a way that B and its tangent line (which is displaced from its original position) may occupy their previous position. This gives us the broken curve $O_{1}B$. $O_{1}O_{2}$ is : $O$'s displacement relatively to $B$ and $B'$s tangent. Now the pier being immovable $O_{1}B$ (right line) = $OB$; i.e., the $X$ projection (or $Z$) of $O_{1}$ upon $OB$ (taken as an axis of $X$) is zero compared with its $\Delta y$. Hence as one condition to fix the special equilibrium polygon for a given loading we have (from § 373) $$\int_{a}^{b} \left( Mydx + EI \right) = 0 \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (1)$$ The other two are that the must pass through $O$. (2) special equilibrium polygon is " " " " " B. (3) Class D. Rib with Fixed Ends and no hinges, i.e., continuous. Piers immovable. The ends may be fixed by being inserted, or built, in the masonry, or by being fastened to large plates which are bolted to the piers. [The St. Louis Bridge and that at Coblenz over the Rhine are of this class.] Fig. 423. In this class there being no hinges we 460 MECHANICS OF ENGINEERING. have no point given in advance through which the special equilibrium polygon must pass. However, since $O'$ displacement relatively (and absolutely) to $B$ and $B'$ tangent is zero, both $\Delta x$ and $\Delta y$ [see § 373] = 0. Also the tangent-lines both at $O$ and $B$ being fixed in direction, the angle between them is the same under loading, or change of temperature, when the rib was first placed in position under stress and at a definite temperature. Hence the conditions for locating the special equilibrium polygon are $$\oint \frac{Mds}{EI} = 0; \oint \frac{Myds}{EI} = 0; \oint \frac{Mdx}{EI} = 0.$$ In the figure the imaginary rigid prolongations at the ends are shown [see § 366]. Other designs than those mentioned are practicable (such as: one end fixed, the other hinged; both ends fixed and one hinge between, etc.), but are of unusual occurrence. 378a. Rib of Three Hinges. Forces not all Vertical. If the given rib of three hinges upholds a roof, the wind-pressure on which is to be considered as well as the weights of the materials composing the roof-covering, the forces will not all be vertical. To draw the special equil. polygon in order to determine the following construction holds: Re- quired to draw an equilibrium polygon, for any plane system of forces, through three arbitrary points, $A$, $p$ and $B$; Fig. 423a. Find the line of action of $R_1$, the resultant of all the forces occurring between $A$ and $p$; also, A diagram showing a rib with three hinges and a roof covering. The diagram includes labeled points A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. For. em. ABCH-RIBS. 461 that of $R_2$, the resultant of all forces between $p$ and $B$; also the line of action of $R_1$, the resultant of $R_2$ and $R_3$ (see § 328). Join any point $M$ in $R$ with $A$ and also with $B$, and join the intersections $N$ and $O$. Then $AN$ will be the direction of the first segment, $OB$ that of the last, and $NO$ itself is the segment corresponding to $p$ (in the desired polygon) an equilibrium polygon for the given forces. See § 328. If $A N^{\prime} p O B'$ are the corresponding segments (as yet unknowns) of the desired equal polygons, we note that the two triangles $MN'N$ and $MN'O$, having their vertices on three lines which meet in a point (i.e., $R$ meets $R_1$ and $R_2$ in $C'$), are homologous [see Prop. VII. of Introductio to Modern Geometry, in Chauvenet's Geometry,] and that . . . the three intersections of their corresponding sides must lie on the same straight line. Of these intersections we already have $A$ and $B$, while the third must be at $C$, found at the intersection of $AH$ and $NO$. Hence by connecting $C$ and $p$, we determine $N'$ and $O'$. Joining $N'A$ and $O'B$, we find that the third diagonal will be to $N'A$, while the last ray will be to $O'B$, and thus the pole of that diagram can easily be found and the corresponding equilibrium polygon, beginning at $A$, will pass through $p$ and $B$. (This general case includes those of §§ 341 and 342.) 379. Arch-Rib of two Hinges; by Prof. Eddy's Method.* [It is understood that the hinges are at the ends.] Required the location of the special equilibrium polygon. We here suppose the rib homogeneous (i.e., the modulus of Elasticity constant throughout); that it is a "cylindrical prism" (i.e., that the moment of inertia I of the cross-section is constant), that the piers are on a level, and that the rib-curve is symmetrical about a vertical line. Fig. 424. For each point m of the rib curve we have an x and y (both known, being the co-ordinates of the point), and also a z (intercept between rib and special equilibrium polygon) and a $\alpha$ (intercept) A diagram showing a curved rib with points labeled A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. *P. 55 of Prof. Eddy's book; see reference in preface of this work. 462 MECHANICS OF ENGINEERING. between the spec. eq. pol. and the axis $X$ (which is $OB$). The first condition given in § 378 for Class C may be transformed as follows, remembering [§ 367 eq. (3)] that $M = Hs$ at any point $m$ of the rib (and that $ET$ is constant): $$\frac{1}{EI} \int_a^b Myds = 0, \text{i.e., } \frac{H}{EI} \int_a^b yds = 0$$ but $$z = y - z' \quad \therefore \int_a^b (y - z')yds = 0; \text{i.e., } \int_a^b yds = \int_a^b y'ds.$$ (1) In practical graphics we cannot deal with infinitesimals; hence we must substitute $\Delta s$ a small finite portion of the rib-curve for $ds$; eq. (1) now reads $\Sigma_y yy ds = \Sigma_y^* yz'ds.$ But if we take all the $\Delta s$ equal, $\Delta s$ is a common factor and cancels out, leaving as a final form for eq. (1) $$\Sigma_y^* (yy) = \Sigma_y^* (yz')$$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The other two conditions are that the special equilibrium polygon begins at $O$ and ends at $B$. (The subdivision of the rib-curve into an even number of equal $\Delta s$ will be observed in all problems henceforth.) 379a. Detail of the Construction. Given the arch-rib $O R_1$, Fig. 425, with specified loading. Divide the curve into Diagram showing a curved line with points labeled O, A, B, C, D, E, F, G, H, I. Fig. 425 ARCH RIBS. 463 eight equal $ds$'s and draw a vertical through the middle of each. Let the loads borne by the respective $ds$'s be $P_1$, $P_2$, etc., and with them form a vertical lead-line $dC$ to some convenient scale. With any convenient pole $O''$ draw a trial force diagram $O''dC$ and a corresponding trial equilibrium polygon $FG$, beginning at any point in the vertical $F$. Its ordinates $z'', z''$, etc., are proportional to the loads $P_1$, $P_2$, etc., still pole sought (whose abscissa line is $OD$ [§ 374a]). Next let it determine $n'$ (see § 374a). We know that $OB$ is the "abutment-line" of the required special polygon, and that its pole must lie on a horizontal through $n'$. It remains to determine its $H_r$ or pole distance, by equation (1)' just given, viz.: $\Sigma_1^r(yg)=\Sigma_1^r(yg')$. First by § 375 find the value of the summation $\Sigma_1^r(yg)$, which, from symmetry, we may write $2\Sigma_1^r(yg)=2y_1y_2+2y_2y_3+2y_3y_4$. Hence, Fig. 426, we obtain $$\Sigma_1^r(yg)=H_rk_r \quad (H_k)$$ Next, also by § 375, see Fig. 427, using the same pole distance $H_r$ as in Fig. 426, we find $$\Sigma_1^r(yg')=H_rk'_r \quad (i.e.,)$$ $$y_{r''}+y_{r'''}+y_{r''''}+y_{r'''''}=H_rk'_r$$ Again, since $\Sigma_1^r(yg')=y_{r''}, y_{r'''}$, which from symmetry (of rib) $$=y_{r''}+y_{r'''}-y_{r''}+y_{r'''}$$ we obtain, Fig. 428, $$\Sigma_1^r(yg')=H_rk'_r \quad (same H_r),$$ and... $$\Sigma_1^r(yg')=H_r(k''+k'').$$ If now we find that $k''+k'=""2k$. 464 MECHANICS OF ENGINEERING. the condition $2y(y) = 2y(y')$ is satisfied, and the pole distance of our trial polygon in Fig. 425, is also that of the special polygon sought; i.e., the $z'$'s are identical in value with the $x'$'s of Fig. 424. In general, of course, we do not find that $k'' + k'' = 2k$. Hence the $z''$'s must all be increased in the ratio $\frac{1}{2} (k'' - k'')$ to become equal to the $x'$'s. That is, the pole distance $H$ of the spec. equil. polygon must be $$H = \frac{k'' + k'''}{2k} H'' \quad \text{(in which } H'' \text{ is the pole distance of the trial polygon)} \quad \text{since from §399 the ordinates of two equilibrium polygons (for the same loads) are inversely as their pole distances. Having thus found the I of the special polygon, knowing that the pole must lie on the horizontal through } x', \text{ Fig. 425, it is easily drawn, beginning at } O. \text{ As a check, it should pass through } B.$$ For its utility see § 367, but it is to be remembered that the stresses as thus found in the different parts of the rib under a given loading, must afterwards be combined with those resulting from change of temperature and the shortening of the rib axis due to the tangential thrusts, before the actual stress can be declared in any part. (Note.—If the "moment of inertia," $I$, of the rib-section is different at different sections, i.e., if $I$ is variable, for eq. (1)' we may write $\frac{x_n}{n}(y_n^{'}) = \frac{x_n}{n}(y_n^{''}) \dots (1)$ (where $n = \frac{I}{I_s}$, $I_s$ being the moment of inertia of a particular section taken as a standard and $I$ that at any section of rib) and in Fig. 426, use the $\frac{x_n}{n}$ of each $J_s$ instead of $y$ in the vertical "load-line," and $\frac{x_n}{n}$ for $z''$ in Figs. 427 and 428). A stylized letter 'C' or a similar shape. ARCH-1138S. 465 380. Arch Rib of Fixed Ends and no Hinges -Example of Class D. Prof. Eddy's Method.* As before, $E$ and $I$ are constant along the rib Piers immovable. Rib curve symmetrical about a vertical line. Fig. 429 shows such a rib under any loading. Its span is $OB$, which is taken as an axis $X$. The co-ordinates of any point $m'$ of the rib curve are $x$ and $y$, and $z$ is the vertical intercept between $m'$ and the special equilibrium polygon (as yet unknown, but to be constructed). Prof. Eddy's method will now be given for finding the sp. A diagram showing a rib with a vertical line through its center, labeled "OA". The diagram includes various lines and points labeled with letters and numbers. Fig. 429. The three conditions it must satisfy (see § 378, Class D, remembering that $E$ and $I$ are constant and that $M = I \frac{d^2}{dx^2}$) are $$\int_0^b x dx = 0; \quad \int_0^b y dx = 0; \quad \int_0^b z dx = 0 \quad . . . \quad (1)$$ Now suppose the auxiliary reference line (straight) $\overline{OM}$ to have been drawn satisfying the requirements, with respect to the rib curve that $$\int_0^b x' dx = 0; \quad \int_0^b y' dx = 0 \quad . . . \quad (2)$$ in which $x'$ is the vertical distance of any point $m'$ from $\overline{OM}$ and $x$ the abscissa of $m'$ from $O$. From Fig. 429, letting $x'$ denote the vertical intercept (corresponding to any $m'$) between the spec. polygon and the auxiliary line $\overline{OM}$, we have $z = x' - x$, hence the three conditions in (1) become $$\int_0^b (x' - x') dx = 0; \quad i.e., \text{ see eq. (2)} \quad \int_0^b x' dx = 0 \quad . . . \quad (3)$$ * P. 14 of Prof. Eddy's book ; see reference in preface of this work. 466 MECHANICS OF ENGINEERING. $$\int_{a}^{x}(x'-x')dx=0;\mathrm{~i.e.,~see~eqs.~(2)~}\int_{a}^{x}xx'ds=0\ \ (4)$$ and $$\int_{a}^{y}(y'-y')dy=0,\ \ \mathrm{by~transposing~position~}\int_{a}^{x}yx'ds=\int_{a}^{y'}ds\ \ (5)\ \ \mathrm{provided~vm~has~been~located~as~prescribed.}$$ For graphical purposes, having subdivided the rive curve into an even number of small equal $J's$, and drawn a vertical through the middle of each, we first, by § 377, locate vm to satisfy the conditions $$\Sigma y'(x')=0\ \mathrm{and}\ \Sigma x(x')=0\ \ \ \ (6)$$ (see eq. (2); the $J's$ cancel out); and then locate the special equilibrium polygon, with vm as a reference-line, by making it satisfy the conditions. $$\Sigma y'(x')=0\ .\ (7);\ \Sigma x(x')=0\ .\ (8);\ \Sigma(y'y')=\Sigma(y'x')\ .\ (9)$$ (obtained from eqs. (3), (4), (5) by putting $da=d_x$, and cancelling). Conditions (7) and (8) may be satisfied by an infinite number of polygons drawn to the given loading. Any one of these being drawn, as a trial polygon, we determine for it the value of the sum $\Sigma(y'y')$ by § 375, and compare it with the value of the sum $\Sigma(y'x')$ which is independent of the special loading, and is thus constant. It must be understood that the quantities (lengths) $x$, $y$, $z$, and $r'$, here dealt with are those pertaining to the verticals drawn through the middles of the respective $J's$, which must be sufficiently numerous to obtain a close result, and not to the verticals in which the loads act, necessarily, since these latter may be few or many according to circumstances, see Fig. 429). If these sums are not equal, the pole distance of the trial equi. polygon must be altered in the proper ratio (and thus change the $x'$ in the inverse ratio) necessary to make both sums equal and thus satisfy conditions (9). The alteration of the $x'$s, all in the same ratio, will ARCH-RIBS. 467 not interfere with conditions (7) and (8) which are already satisfied. 381. Detail of Construction of Last Problem. Symmetrical Arch-Rib of Fixed Ends.—As an example take a span of the St. Louis Bridge (assuming $I$ constant) with "live load" coverring the half span on the left, Fig. 450, where the vertical A diagram showing a symmetrical arch-rib of fixed ends, with a live load covering the half span on the left side. Fig. 450. scale is much exaggerated for the sake of distinctness*. Divide into eight equal $d's$. (In an actual example sixteen or twenty should be taken.) Draw a vertical at each end of the $d$'s. The moment of inertia, however, is not strictly constant, the portion near each pier, of a length equal to one eighth of the span, having a value of $I$ one-half greater than that of the remainder of the arc. * Each arch-rib of the St. Louis bridge is a bulb up or trussed rib of steel about an 8' span and 38 ft. rise, in the form of a segment of a circle. Its moment of inertia, however, is not strictly constant, the portions near each pier, of a length equal to one eighth of the span, having a value of $I$ one-half greater than that of the remainder of the arc. 468 MECHANICS OF ENGINEERING. middle of each $J_0$. $P_1$, etc., are the loads coming upon the respective $J_0$'s. First, to locate $v_m$, by eq. (6); from symmetry it must be horizontal. Draw a trial $v_m$ (not shown in the figure) and if the ($+r'$)'s exceed the ($-r'$)'s by an amount $n'$, the true $v_m$ will lie a height $\frac{1}{n}z'_m$ above the trial $v_m$ (or below, if vice versa); $n = n$ the number of $J_0$'s. Now lay off the load-line on the right (to scale), take as convenient trial pole $O''$ and draw a corresponding trial equil. polygon $F''G''$ by § 377, locate a straight line $v''m$ so as to make $\Sigma x''(x'')=0$ and $\Sigma y''(y'')=0$ (see Note (b) of § 377). [We might now redraw $F''G''$ in such a way as to bring $v''m''$ into a horizontal position, thus: first determine a point $n''$ on the load-line by drawing $O''n''$ to $v''m''$ take a new pole on a horizontal through $n''$, with the same $H''$, and draw a corresponding equil. polygon; in the latter $v''m''$ would be horizontal. We might also shift this new trial line to coincide with the original one when they coincide. It would satisfy conditions (7) and (8), having the same $x''$'s as the first trial polygon; but to satisfy condition (9) it must have its $x''$'s altered in a certain ratio, which we must now find. But we can deal with the individual $x''$'s just as well in their present positions in Fig. 430.] The points $E$ and $L$ in $v_m$, vertically over $E''$ and $L''$ in $v''m$, are now fixed; and are the intersections of the special polygon required, with $v_m$. The lines drawn between these points and the trial equilibrium polygon have been called $x''$ instead of $x'$; they are proportional to the respective $x''$'s of the required special polygon. The next step is to find in what ratio the ($x''$)'s need to be altered (or $H''$ altered in inverse ratio) in order to become the ($x'$)'s; i.e., in order to fulfil condition (9), viz: A diagram showing various lines and points labeled with symbols like J0, P1, v_m, z'_m, O'', F'', G'', n'', E, L, E'', L'', x'', y'', H'', and other geometric elements. 409 AECB-RIBS. (9) This may be done precisely as for the rib with two hinges, but the negative ($r$')s must be properly considered (§ 375) See Fig. 431 for the detail. Negative $r$'s or $r$'s point upward. From Fig. 431a $\Sigma(yz') = H_k$ from symmetry $\Sigma(yz') = 2H_k$ From Fig. 431b we have $\Sigma(yz') = H_k$ and from Fig. 431c $\Sigma(yz') = H_k$ (The same pole distance $H_c$ is taken in all these constructions.) $\Sigma(yz'') = H_c(k + k)$. If, then, $H_c(k + k) = 2H_c$ condition (9) is satisfied by the $y''$'s. If not the true pole distance for the special equil. polygon of Fig. 430 will be $$H = \frac{k + k}{2k} H''$$ With this pole distance and a pole in the horizontal through $y''$ (Fig. 430) the force diagram may be completed for the required special polygon ; and this latter may be constructed for any given value of $y''$. In the polygon in which through it draw a segment $l$ to the proper ray of the force diagram. In our present figure (430) this "proper ray" would be the ray joining the pole with the point of meeting of $P_1$ and $P_2$ on the load-line. Having this one seg- 470 MECHANICS OF ENGINEERING. most of the special polygon the others are added in an obvious manner, and thus the whole polygon completed. It should pass through $L$, but not $O$ and $B$. For another loading a different special equil. polygon would result, and in each case we may obtain the thrust, shear, and moment of stress couple for any cross-section of the rib, by § 367. To the stresses computed from these, should be added (algebraically) those occasioned by change of temperature and by shortening of the rib as occasioned by the action of the load, and also "temperature stresses," and stresses due to rib-shortening, will be considered in a subsequent paragraph. They have no existence for an arch-rib of three hinges. [Note.—If the moment of inertia of the rib is variable we put $\frac{z'}{n}$ for $z'$ and $\frac{z''}{n}$ for $z''$ in equation (6), (7), (8), and (9), $s$ having the meaning given in the Note in § 379 a, which see; and proceed accordingly.] 381a. Exaggeration of Vertical Dimensions of Both Space and Force Diagrams.—In case, as often happens, the axis of the given rib is quite a flat curve, it is more accurate (for finding $M$) to proceed as follows: After drawing the space diagram with its true proportions and passing a vertical through the middle of each of the equal $de's$, compute the ordinate ($y$) of each of these middle points from the equation of the curve, and multiply each $y$ by four (say). These quadruple ordinates are then laid off from the span upward, each in its proper vertical. Also multiply each load, of the given loading, by four, and then with these quadruple loads and quadruple ordinates, and the upper extremities of the latter as points in an exaggerated curve, proceed to construct a special equilibrium polygon, and thus correct force diagram by the proper method ( for Class B, C or D, as the case may be) for this exaggerated rib-curve. The moment, $Hs$, thus found for any section of the ex- ARCH-RIBS. 471 aggregated rib-curve, is to be divided by four to obtain the moment in the real rib, in the same vertical line. To find the thrust and shear, however, for sections of the real rib, besides employing tangents and normals of the real rib we must draw, and use, another force diagram, obtained from the one already drawn (for the exaggerated rib) by reducing its vertical dimensions (only), in the ratio of four to one. (Of course, any other convenient number besides four, may be adopted throughout.) 382 Stress Diagrams.—Take an arch-rib of Class $D$, § 378, i.e., of fixed ends, and suppose that for a given loading (including its own weight) the special equil polygon and its force diagram have been drawn (§ 381). The diagram re- quired to indicate graphically the variation of the three stress-elements for any section of the rib, viz., the thrust, shear, and mom. of stress-couple. $I$ is constant. If at any point $m$ of the rib a section is made, then the stresses in that section are classified into three sets (Fig. 432). (See §§ 255 and 307) and from § 307 eq. (3) we see that the vertical lines between both sets of ribs, for the special equil polygon being proportional to the products $P_0^2$ for moments of the stress-couples in the corresponding sections form a moment-diagram, on inspection of which we can trace the change in this moment, $H = \frac{Py}{e}$, and hence the variation of the stress per square inch, $\rho_b$ (as due to stress couple alone) in the outermost fibre of any section (tension or compression) at distance $e$ from the gravity axis of the section), from section to section along the rib. By drawing through 0 lines On' and Ot' parallel respectively to the tangent and normal at any point $m$ of the rib axis [see Fig. 433] and projecting upon them, in turn, the proper ray ($R$, in Fig. 433) [see eqs. 1 and 2 of § 307] Fig. 432 Fig. 433 472 MECHANICS OF ENGINEERING. we obtain the values of the thrust and shear for the section at $m$. When found in this way for a number of points along the rib their values may be laid off as vertical lines from a horizontal axis, in the verticals containing the respective points, and thus a thrust diagram and a shear diagram may be formed, as constructed in Fig. 433. Notice that where the moment is a maximum or minimum the shear changes sign (compare § 240), either gradually or A diagram showing a curved line representing a rib with various points labeled P, Q, R, S, T, U, V, W, X, Y, Z. The curve is labeled "THROTTLE" and "WINDS". There are also labels for "MOMENTS" and "THRUST". Fig. 433. suddenly, according as the max. or min. occurs between two loads or in passing a load ; see $m'$, e. g. Also it is evident, from the geometrical relations involved, that at those points of the rib where the tangent-line is parallel to the "proper ray" of the force diagram, the thrust is a maximum (a local maximum) the moment (of ARCH-RIBS 473 stress couple) is either a maximum or a minimum and the shear is zero. From the moment, $Hs = \frac{p_0L}{e} p_t = \frac{Hs}{I}$ may be computed. From the thrust = $Fp$, $p_t = \frac{\text{thrust}}{F}$, ($F$ = area of cross-section) may be computed. Hence the greatest compression per sq. inch $(p_0 + p_t)$ may be found in each section. A separate stress-diagram might be constructed for this quantity $(p_0 + p_t)$. Its max. value (after adding the stress due to change of temperature, or to shortening, for ribs of less than three hinges), wherever it occurs in the rib, will make the rib safe by proper designing of the rib. The maximum shear $\sigma_c$ can be used as in §256 to determine thickness of web, if the section is I-shaped, or box-shaped. See § 295. 383. Temperature Stresses.—In an ordinary bridge truss and straight horizontal girders, free to expand or contract longitudinally, and in Classes A and B of § 378 of arch-ribs, there are no stresses induced by change of temperature; for the form of the beam or truss is under no constraint from the manner of support; but with the arch-rib of two hinges (hinged ends, Class C) and of fixed ends (Class D) having immovable pieces which constrain the distance between them. When the temperature changes, stresses called “temperature stresses” are induced in the rib whenever the temperature, $t$, is not the same as that, $t_o$, when the rib was put in place. These may be determined, as follows, as if they were the only ones, and then combined, algebraically, with those due to the loading. 384. Temperature Stresses in the Arch-Rib of Hinged Ends—(Class C, § 378.) Fig. 434. Let $E$ and $I$ be constant, with 474 MECHANICS OF ENGINEERING. g. H other postulates as in § 373. Let $t_s$ = temperature of erection, and $t_c$ = any other temperature; also let $l$ = length of span = $OB$ (invariable) and $\gamma$ = co-efficient of expansion of the material of the curved beam or rib (see § 395). At the pressure $t$ there must be a horizontal reaction $H$ at each hinge to prevent expansion into the form $OB$ (dotted curve), which is the form natural to the rib for tenperature $t$ and without constraint. We may consider the actual form $OB$ as having resulted from the unstrained form $OB'$ by displacing $O'$ to $O$, i.e., producing a horizontal displacement $O'O' = l(t - t_c)$. But $(O'O')^2 = (\frac{1}{E} \cdot M)^2 + (\frac{1}{I} \cdot H)^2$ (§ 373 and 374). (N.B. $F'$s tangent has moved, but this does not affect $Jc$, if the axis $x$ is horizontal, as here, coinciding with the span;) and the ordinate $y$ of any point $m$ of the rib is identical with its $x$ or intercept between it and the spec. equil. polygon, which here consists of one segment only, viz.: $OB$. Its force diagram consists of a single ray $O', m'$; see Fig. 434. Now (§ 373) $$dx = \frac{1}{EI} \int yds; \quad M = Hx \quad \text{is in this case, } Hy$$ :. $(t - t_c) \gamma = \frac{H}{EI} \int yds; \quad \text{hence for graphics, and equal } Jc's, we have}$$ $$\frac{EI}{(t - t_c)} H dx = \frac{1}{2} y^2 \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)$$ From eq. (1) we determine $H$, having divided the rib-curve into from twelve to twenty equal parts each called $dx$. For instance, for wrought iron, $t$ and $t_c$ being expressed in Fahrenheit degrees, $\gamma = 0.000066$. If $E$ is expressed in lbs. per square inch, all linear quantities should be in inches and $H$ will be obtained in pounds. $\gamma y^2$ may be obtained by § 375, or may be computed. $H$ being known, we find the moment of stress-couple = $Hy$, ARCH-RIBS. 475 at any section, while the thrust and shear at that section are the projections of $H$, i.e., of $qN$ upon the tangent and normal. The stresses due to these may then be determined in any section, as already so frequently explained, and then combined with those due to loading. 385. Temperature Stresses in the Arch-Ribs with Fixed Ends—See Fig. 435. (Same postulates as to symmetry, $E$ and $I$ constant, etc., as in § 380.) $t$ and $\ell$, have the same meaning as in § 384. Here, as before, we consider the rib to have reached its actual form under temperature $t$ by having had its length shortened from the length natural to temp. $t$, viz.: $O'B'$, to the actual length $OB$, which the immovable piers compel it to assume. But here, since the tangents at $O$ and $B$ are to be the same in direction under constraint as before, the two forces $H$, representing the action of the piers on the rib, must be considered as acting on imaginary rigid prolongations at an unknown distance $d$ above the span. To find $H$ and $d$ we need two equations. From § 373 we have, since $M = Hx = H(y - d)$, $$dx, \text{ i.e., } O'T + BE', \text{ i.e., } H(t - t_0) = EI\int_{y-d}^{y} dyds \quad (2)$$ or, graphically, with equal $ds's$ $$EI(t - t_0)y = Hds\left[\Sigma_y y^2 - d\Sigma_y y\right] \quad . . . \quad (3)$$ Also, since there has been no change in the angle between end-tangents, we must have, from § 374, $$\frac{1}{EI}\int dsds = 0; \text{ i.e., } \frac{H}{EI}\int dsds = 0; \text{ i.e., } f_y(y - d)ds = 0$$ 476 MECHANICS OF ENGINEERING. or for graphics, with equal $Ae's$, $Xy = nd \ldots$ (4) in which $n$ denotes the number of $Ae's$. From (4) we determine $d$, and then from (5) can compute $H$. Drawing the horizontal $FG$, it is the special equilibrium polygon (of but one segment) and the moment of the stress-couple at any section is $Hd$, while the thrust and shear are the projections of $H=0.5a$ on the tangent and normal respectively of any point $m$ of rib. For example, a rib of 550 feet, of the St. Louis Bridge, having a rise of 85 feet and fixed at the ends, the force $H$ of Fig. 435 is = 108 tons, when the temperature is 80° Fahr. higher than the temp. of erection, and the enforced span is $3\frac{1}{2}$ inches shorter than the span natural to that higher temperature. Evidently, if the actual temperature is lower than that of erection, $H$ must act in a direction opposite to that of Figs. 435 and 434, and the "thrust" in any section will be negative, i.e., a pull. 386. Stresses Due to Rib-Shortening.—In § 389, Fig. 407, the shortening of the element $AE$ to a length $AE'$ due to the uniformly distributed thrust, $pAE'$, was neglected as producing indirectly a change of curvature and form in the rib axis; but such will be the case if the rib has less than three hinges. This change in the length of the different portions of the rib may be made up by a change due to a change of temperature. For example, from § 199 we see that a thrust of 50 tons coming upon a sectional area of $F = 10$ sq. inches in an iron rib, whose material has a modulus of elasticity $E = 30,000,000$ lbs. per sq. inch, and a coefficient of expansion $\gamma = -0.000066$ per degree Fahrenheit, produces a shortening equal to that due to a fall of temperature $(t_2 - t)$ derived as follows: (See § 199) $$\left(t_2 - t\right) = \frac{P}{FEg} = \frac{100,000}{10 \times 30,000,000 \times 0.000066} = 5^{\circ}\text{Fahrenheit}$$ Practically, then, since most metal arch bridges of classes C and D are rather flat in curvature, and the thrust ARCH-RIBS. 477 due to ordinary modes of loading do not vary more than 20 or 30 per cent from each other along the rib, an imaginary fall of temperature corresponding to an average thrust in any case of loading may be made the basis of a constructional calculation. In 1858 (according as the units are *hinged*, or *fixed*), from which the stresses due to thrusts and stress-couple moments, may be derived to be combined with those previously obtained for loading and for change of temperature. **267. Resumé.** It is now seen how the stresses per square inch, both shear and compression (or tension) may be obtained by parts, and a section of a curved beam of the kinds described, by combining the results due to the three separate causes, viz.: the load, change of temperature, and rib-shortening caused by the thrusts due to the load (the latter agglomeration, however, coming into consideration only in classes $C$ and $D$, see § 375). That is, in any cross-section, the stress in the outer fibre is, [letting $T_1^{'}, T_2^{'}, T_3^{'},$ denote the thrusts due to the three causes, respectively, above mentioned; $(Hb)^{'}, (Hc)^{'},$ $(Ha)^{'},$ the moments] $$T_i^{'\pm} + \frac{T_i}{F} + T_i^{''\pm} + \left[ \left( Hb \right)^{\prime\prime} \pm \left( Hc \right)^{\prime\prime} \pm \left( Ha \right)^{\prime\prime} \right]$$ (1) i.e., lbs. per sq. inch compression (if those units are used). The double signs provide for the cases where the stresses in the outer fibre, due to thrusts and temperature changes, are opposite. Fig. 436 shows the meaning of $\epsilon$ (the same used heretofore) $I$ in the moment of inertia of the section about the gravity axis (horizontal) $C$. $F =$ area of cross-section. $\left[ e = e' \right]$ cross section symmetrical about $C$. For a given loading we may find the maximum stress in a given rib, or design the rib so that this maximum stress shall be safe for the material employed. Similarly, the resultant shear (total, not A diagram showing a curved beam with various forces acting on it. Fig. 436. 478 MECHANICS OF ENGINEERING. per sq. inch) = $J^{\prime} + J^{\prime\prime} + J^{\prime\prime\prime}$ is obtained for any section to compute a proper thickness of web, spacing of riveta, etc. 388 The Arch-Tree, or braced arch. An open-work truss, if of homogeneous design from end to end, may be treated as a beam of constant section and constant moment of inertia, and if curved, like the St Louis Bridge and the Coblenz Bridge (see § 378, Class D), may be treated as an arch-rib. The moment of Inertia may be taken as $$I = 2F_i \left(\frac{h}{2}\right)^2$$ where $F_i$ is the sectional area of one of the pieces i to the curved axis midway between them, Fig. 437, and $h$ = distance between them. Treating this curved axis as an arch-rib, in the usual way (see preceding articles), we obtain the spec. equil. pol. and its force diagram for given loading. Any plane $\tau$ to the rib-axis, where it crosses the middle m of a "web-member", cuts three pieces, $A$, $B$ and $C$, the total com- The St. Louis Bridge is not strictly of constant moment of inertia, being somewhat strengthened near each pier. Fig. 436. Fig. 437. ARCH-RIBS. 479 pressions (or tensions) in which are thus found : For the point $m$, of rib-axis, there is a certain moment = $H_0$ at thrust = $T_0$, and a shear = $J$, obtained as previously explained. Hence, by (1), $P' + P'' = H_0 \cos \beta$ (1) and thus determine whether $P'$ is tension or compression ; then putting $P' + P'' \pm P \cos \beta = T_0$. . . . . . 2 (in which $P$ is taken with a plus sign if a compression, and minus if tension), and $$\left(P'-P''\right)_B = H_0$$ (3) we compute $P'$ and $P''$, which are assumed to be both compressions here. $\beta$ is the angle between the web member and the tangent to rib-axis at $m$, the middle of the piso. See Fig. 406, as an explanation of the method just adopted. HORIZONTAL STRAIGHT GIRDSERS. 388. Ends Free to Turn.—This corresponds to an arch-rib with hinged ends, but it must be understood that there is no hindrance to horizontal motion. (Fig. 430.) In Diagram showing a straight beam slightly bent under vertical forces only (as in this case with no horizontal constraint). Fig. 430. treating a straight beam, slightly bent under vertical forces only (as in this case with no horizontal constraint), as 480 MECHANICS OF ENGINEERING. particular case of an arch-rib, it is evident that since the pole distance must be zero, the special equil. polygon will have all its segments vertical, and the corresponding force diagram reduces to a single vertical line (the load line). The first and last segments must pass through $A$ and $B$ (points of no moment) respectively, but being vertical will not intersect $P_1$ and $P_2$; i.e., the remainder of the special equilibrium polygon lies at an infinite distance above the span $AB$. Hence the actual spec. equil. pol. is useless. However, knowing that the shear, $J$, and the moments $M$ of stress on any section of the beam are proportional to any section $m$ (Fig. 439) which we wish to determine (since there is no thrust along the beam), and knowing that an imaginary force $H''$, applied horizontally at each end of the beam, would have no influence in determining the shear and moment at $m$ as due to the new system of forces, we may therefore obtain the shears and moments graphically from this new system (viz.: the loads $P_n$ etc., the vertical reactions $V$ and $V_n$, and the two equal and opposite $H''$'s). [Evidently, since $m$ has no moment about the horizontal axis (or gross axis) because of the moment at $m$ will be unaffected by it; and since $H''$ has no component $\uparrow$ to the beam at $m$, the shear at $m$ is the same in the new system of forces, as in the old, before the introduction of the $H''$'s.] Hence, lay off the load-line 1...2...3, Fig. 439, and construct an equil. polyg., which shall pass through $A$ and $B$ and have any convenient arbitrary $H''$ (force) as a pole distance. Draw a line parallel to this pole distance through $O$, using the auxiliary polygon $\overline{OA}H'$, to make $\overline{OA}$ (arbitrary), and drawing $\overline{O'}n'$ to $\overline{AB}$'. Taking $\overline{O'}n'$ on a horizontal through $\overline{n'}$, making $\overline{O'n'}=\overline{H''}$', we complete the force diagram, and equil. pol. $\overline{AB}$. Then, $\overline{s}$ being the vertical intercept between $m$ and the equil. polygon, we have: Moment at $m=\overline{M}=\overline{H''}\cdot\overline{s}$ (or $\overline{H''}$ also), and shear at $m$, or $\overline{J}=\overline{s}\cdot\overline{n'}$, i.e., = projection of the proper ray $\overline{R_n}$ or $\overline{O''}-\overline{2}$, upon the vertical through $m$. Similarly we obtain $\overline{M}$ and $\overline{J}$ at any other section for the given load. (See ARCH-RIBS; SPECIAL CASE ; STRAIGHT. 481 §§ 339, 337 and 367). The moment of inertia need not be constant in this case. 390. Straight Horizontal Prismatic Girder of Fixed Ends at Same Level—No horizontal constraint, hence no thrust. I constant. Ends at same level, with end-tangents horizontal. We may consider the whole beam free (cutting close to the walls) putting in the unknown upward shears $J_1$ and $J_2$. Diagram showing a straight horizontal prismatic girder with fixed ends at the same level. Fig. 465 and the two stress couples of unknown moments $M_1$ and $M_2$ at these end sections. Also, as in § 388, an arbitrary $H''$ horizontal and in line of beam at each extremity. Now (See Fig. 33) the couple at $O$ and the force $H''$ are equivalent to a single horizontal $H'$ at an unknown vertical distance $c$ below $O$; similarly at the right hand end. The special polygon $FG$ is to be determined for this new system, since the moment under this new system will be the same at any section under this new system as under the real system. The couple $M_1d_1$ is determined as follows: Since the end-tangents are fixed, $V'M_1d_1 = V'_{\text{end}}d_1 = 0$ and since $$ \begin{align*} & \text{at } O: \\ & M_1d_1 + H'd_1 = 0 \\ & \text{at } d_2: \\ & M_2d_2 - H'd_2 = 0 \end{align*} $$ 482 MECHANICS OF ENGINEERING. $O$a's displacement relatively to $B$a's tangent is zero we have $\Sigma M_{a}d_{a}=0\ldots \Sigma F^{a}m_{a}d_{a}=0\ldots \Sigma E_{a}d_{a}=0$. See § 374. Hence for equal $d_{a}$s, $\Sigma q_{a}=0$ and $\Sigma x_{a}=0$. Now for any pole $O^{''}$ draw an equil. pole $F^{''}G^{''}$ and in it (by § 377; see Note) locate $v^{''}m^{''}$ so as to make $\Sigma y^{''}=0$ and $\Sigma L^{''}=0$. Draw verticals through the intersections $E^{''}$ and $L^{''}$, to determine $E$ and $L$ on the beam, these are the points of inflection (i.e., of zero moment), and are points in the required special polygon $FG$. Draw $O^{''}v^{''}$ to $v^{''}m^{''}$ to fix $v^{''}$. Take a pole $O^{'}$ on the horizontal line of $O^{''}v^{''}$ making $y^{'}z^{'}=H^{'}$ (arbitrary), draw the force diagram $O^{'}1234$ and a corresponding equilibrium polygon beginning at $E$. It should cut $I$, and will fulfil the two requirements $\Sigma y=0$ and $\Sigma z=0$, with reference to the axis of the beam $O^{'}B^{'}$. The moment of the stress-couple at any section $m$ will be $M=I^{'P}u'$, and the shear $J=$ the projection of the "proper ray" of the force diagram $O^{'}1\ldots 2$, etc., upon the vertical (not in the trial diagram $O^{'}1\ldots 2$, etc.). As far as the moment is concerned the whole polygon is a "proper" one, but for the special polygon $FG$ i.e., $M=M^{''}v^{''}$ as well as $H'=H'^{'}$, $H'^{'}$ being the pole-distance of $O^{'}$; but for the shear we must use the rays of the final and not the trial diagram. The peculiarity of this treatment of straight beams, considered as a particular case of curved beams, consists in the substitution of an imaginary system of forces involving the two equal and opposite, and arbitrary $H'a$, for the real system in which there is no horizontal force and consequently no special "equilibrium polygon", and thus determining all that is desired, i.e., the moment and shear at any section. In the polygon $FG$ the student will recognize the "moment-diagram" of the problems in Chaps. III and IV. He will also see why the shear is proportional to the slope $\frac{AM}{AE}$ of the moment curve in those chapters. For example, the "slope" of the second segment of the polygon $FG$, that segment being 1 to $O^{'}2$, is ARCH-RIBS; SPECIAL CASE; STRAIGHT. 483 tang. of angle $2O''n'' = \overline{2n''} + O''n''$ = shear + $H''$ and similarly for any other segment; i.e., the tangent of the inclination of the "moment curve," or line, is proportional to the shear. It is also interesting to notice with the present problem of a straight beam, that in the conditions $\Sigma(xde) = 0$ and $\Sigma(xde)x = 0,$ for locating the polygon $FG$, each de is $\frac{1}{\pi}$ to its $z$, and that consequently each $xde$ is the area of a small vertical strip of area between the beam and the polygon, and $(xde)x$ is the "moment" of this strip of area, about $O'$ the origin of $x$. Hence these conditions imply, first, that the area $EWL$ between the polygon and the axis of the beam on one side is equal to that ($OPE+LBG$) on the other side, and, secondly, that the centre of gravity of $EWL$ lies in the same vertical as that of $OPE+LBG$. Another way of stating the same thing is that, if we join $FG$, the area of the trapezoid $FO'HG$ is equal to that of the figure $FEWLG$, and their centres of gravity lie in the same vertical. A corresponding statement may be made (if we join $F''G''$) for the trapezoid $F''e''m''G''$ and figure $F''B''W''L''G''$. 484 MECHANICS OF ENGINEERING. CHAPTER XIL GRAPHICS OF CONTINUOUS GIRDERS. [MAINDLY DUE TO PROP. MOHR OF AIX-LA-CHAPELLE.] 391. The Elastic Curve of a Horizontal Loaded Beam, Homogeneous and Originally Straight and Prismatic, is an Equilibrium Polygon, whose "load-line" is vertical and consists of the successive products $Mdz$ (treated as if they were loads each applied through the middle of its proper $dz$, and whose "pole distance" is $EL$. Fig 441 (exaggerated). Let $AO$ and $OC$ be any two consecutive equal elements of a very flat-elastic curve (as above described). Prolong $AO$ to cut $NC$. Then from § 231, eq. (7), we have $$\frac{d^2y}{dx} = \frac{(Mdz)}{EI}$$ (1) where $d^2y = \overline{DC}$; and, hence, if a triangle (Fig. 442) $OD'C''$, be formed with $OD'$ to $OD$, $OC''$ to $OC$, and $D'C''$ vertical, while its (horizontal) altitude $C'n$ is made equal, by scale, to $El$ of the beam, then from the similarity of the triangle $ODC$ and $OD'C''$ and the proportion in which we see that $Mdz$ is represented through $Mdz'$ on the same scale by which $O'C''$ represents $El$. It is the moment of A diagram showing a polygonal load line and pole distance EL. A diagram showing a polygonal load line and pole distance EL. A diagram showing a polygonal load line and pole distance EL. A diagram showing a polygonal load line and pole distance EL. CONTINUOUS GIRDER BY GRAPHICS. 483 the stress-couple at the section whose neutral axis is projected in $O$. Similarly, if $M'$ is the moment of the stress-couple at $C$, and we draw $O'F',i$ to $CF$, $F'E$ must represent $M'dx$ (on same scale). It is therefore apparent that the line $AOCF$ bears to the figure $O'F'C'F'$ the same relation which an equilibrium polygon (for vertical forces) does to its force diagram, the "loads" of the force-diagram being the successive values of $Mdx$ laid off to scale, while its "pole-distance" is $EI$ laid off on the same scale. [As if $Mdx$ and $M'dx$ were loads suspended at $O$ and $C$ respectively.] Practically, since any actual elastic curve is very flat, and since a change of pole-distance will change all vertical dimensions of the equilibrium polygon in an inverse equal ratio, we may exaggerate the vertical dimensions of the elastic curve by choosing a pole distance smaller than $EI$ in any convenient ratio, $\alpha$. Any deflection in the elastic curve thus obtained will be greater than its true value in the same ratio $\alpha$. Graphically, in order to draw exaggerated elastic curves according to this principle, we obtain approximate results by dividing the length of the beam into a number of equal $dx's$, draw verticals through the middles of the $dx's$ as "force-verticals," and lay off as a "load-line" to any convenient scale the corresponding values of $Mdx$ in their proper order. The quality of the product $Mdx$ is evidently (length)$^2 \times$ force, and with the foot and pound as units such a product may be called so many (sq. ft.) (lbs.). It will be noticed that these products ($Mdx$) are perpendicular to and may be represented by the lines of those corresponding vertical strips of the "moment-diagram" proper to the case in hand. These strips together make up the "moment-area," as it may be called, lying between the moment curve and its horizontal axis (which is the axis of the beam itself, according to §§ 389 and 390). A diagram showing a continuous girder with a neutral axis projected in O. A diagram showing a stress-couple at C and how it relates to an equilibrium polygon for vertical forces. A diagram showing how to draw exaggerated elastic curves by choosing a pole distance smaller than EI in any convenient ratio $\alpha$. The verticals drawn through the middles of the dx's represent "force-verticals," and the load-line represents the corresponding values of Mdx in their proper order. 486 MECHANICS OF ENGINEERING. § 392. Mohr's Theorem.-The principle of the previous paragraph may therefore be enunciated as follows: That just as the moment curve (of a straight prismatic horizontal beam) between two consecutive supports is an equilibrium polygon for the loading between those supports, so also is the elastic curve itself an equilibrium polygon for the "moment-area" considered as a loading. In dealing with the moment-curves of a single span the pole distance is arbitrary ($\S$ 389 and 390), but the position of the pole $O'$ must be on a fixed line ($\S$ 391), and the location of the moment-curve (equi-pol) relatively to the beam (considered to be still straight for this purpose), depend on whether the beam simply rests on the two supports, without projecting beyond ; or is built in, and at what angles ; or as with a continuous girder, on the inclination of the tangent-lines at the supports, as influenced by the presence of loads on all the spans, and on whether all supports are on the same level or not. For example, in § 389, when a single span, the end of beam being simply supported without overhanging, the pole $O'$ must be on a horizontal through $n'$, and the moment curve must pass through the extremities $A$ and $B$ of the beam, thus giving a "moment-area" lying entirely on one side of the beam (or axis from which the moment ordinates, $z$, are to be measured); whereas, in § 390, also a single span, where the ends of the beam are built in horizontally and at the same level, the pole must be taken on the horizontal through $n$, and the moment-curve $MOLG$ must intersect the base in the points $E$ and $I$ ($E$, $L$, and $n'$ being found as prescribed in that problem), and thus lie partly above and partly below the beam. It will be necessary, later, to distinguish the upper and lower parts of the moment-area as positive and negative. In drawing the equilibrium polygon which constitutes the actual elastic curve, however, and hence making use of the successive small vertical strips of the moment-areas, (when found) as if they were loads, to form a load-line according to a convenient scale, the pole distance is not ar- CONTINUOUS GIRDER BY GRAPHICS. 487 binary but must be = $EI$ on the same scale. Still, since for convenience we must always greatly exaggerate the vertical scale of the elastic curve, we may make the pole distance = $EI + e$ and thus obtain an elastic curve whose vertical dimensions are a times as large as those of the real curve; while the position of the pole will depend in the direction of the tangent lines at the extremities of the span. An example will now be given. 393 Example of an Elastic Curve (Beam Prismatic) Drawn as an Equilibrium Polygon Supporting the Moment-Area as Loading. —Let the beam be simply supported at its extremities (at the same level), and bear a single eccentric load $P$, Fig. 443, its own weight being neglected. The moment-area A diagram showing a triangle ABC with points labeled A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S. Fig. 443 consists of a triangle $ACB$ (see first part of § 260, or use the graphic method of § 389, thus utilizing a force diagram 012), its altitude being the moment represented by the ordinate $CD$ and having a value $\frac{PL}{l}$. Hence the total moment-area = $\frac{PL}{l}$ base $AB \times$ mom. $UD$ i.e., $\frac{PL}{l} \times \frac{1}{2} l = \frac{1}{2} PLd_4$ 488 MECHANICS OF ENGINEERING. Divide $AB$ into (say) eight equal $dx$'s (eight are rather few in practice; sixteen or twenty would be better) and draw a vertical through the middle of each. Note the portion of each of these vertical intercepts between the axis of the beam and the moment-area $ACB$. The products $M dx$ for the different subdivisions are proportional to these intercepts, since all the $dx$'s are equal, and are the respective moment-areas of $ACB$'s. Treating these areas as if they were loads, we lay off the corresponding intercepts (or their halves, or quarters, or other convenient fractional part or multiple), from $E$ downwards to form a vertical "load-line," beginning with the left-hand intercept and continuing in proper order. As to what scale this implies, we determine by dividing the total moment-area thus laid off, viz.: $\frac{1}{2} P l_0$, say in (sq. in.) (lbs.), by the length of $E$ in inches, thus obtaining the number of (sq. in.) (lbs.) which each linear inch of paper represents. On this scale the number of inches of paper required to represent the $EI$ of the beam is so enormous, that in its stead we use the nth portion, $n$ being an arbitrary abstract number of such magnitude as to make $EI + n$ a convenient pole-distance. The proper position of the pole $O'$ on the vertical $TW$, is fixed by the fact that the elastic curve, beginning at $A$, must terminate in $B$, at the same point as $C$. Hence, assuming a trial position for $O'$, and drawing rays in the usual manner (except that, as hereofore, the pole is taken on the right of the "load-line," instead of on the left, so as to make the resulting equilibrium polygon correspond in direction of curvature to the actual elastic curve), we draw the corresponding equilibrium polygon $A''B''$. Determining $n'$ by drawing through $O''$ a line $\parallel$ to the right line $A''O''$, we draw a horizontal through $n'$ to intersect $TW$ at which requires pole. With $O'$ as point, an equilibrium polygon begun at $A'$ will terminate in $B'$ and its vertical ordinates will be a times as great as those of corresponding points on the CONTINUOUS GIRDER BY GRAPHICS. 480 actual elastic curve $AB$. The same relation holds between the tangents of the angle of inclination to the horizontal at corresponding points (i.e., those in same vertical) of the two curves. 394. Numerical Case of Foregoing Example.—With the inch and pound as units, let $P = 120$ lbs., $l_1 = 40$ in., $l_2 = 60$ in., while the prismatic beam is of timber having a modulus of elasticity $E = 2,000,000$ lbs. per sq. inch, and is rectangular in section, being 2 in. wide and 4 in. high, so that its width being placed on the scale of the moment-area-diagram (§ 391) is $\frac{1}{2} \times 2 \times 4 = 4$ square inches. Required the maximum deflection. Adopting 1:20 as the scale for distances (i.e., one linear inch of paper to twenty inches of actual distance) we make the horizontal $AB$ in. long, Fig. 443, and $AD$ in. in., taking a point $C$ at convenience in the vertical through $D$, and joining $AC$ and $CB$, thus determining the moment-diagram for this case. [As to what pole distance, $H$, is implied in this selection, $C_1$ is immaterial in this simple case of a continuous beam having no intermediate supports (corresponding force-diagram at all.) We divide $AB$ into eight equal parts and draw a vertical through the middle of each. The intercepts, in these verticals, between $AB$ and the broken line $ACB$ we lay off from $E$ toward $F$ as prescribed in § 393. (By taking $DC$ small enough the line $EF$ will not be inconveniently long.) Suppose this length $EF$ measures 6.4 inches on the paper (as in the actual draft by the writer). Since it represents a moment of $\frac{1}{2} \times P \times l_1 \times l_2 = \frac{1}{2} \times 120 \times 40 \times 60 = 192,000$ (sq. in.) (lbs.), the scale of our "moment-area-diagram," as we may call it, must be $192,000 + 6.4 = 30,000$ (sq. in.) (lbs.) per linear inch of paper. Now $EI = 21,338,333$ (sq. in.) (lbs.), which on the above scale would be represented by 711 linear inches of paper. With $n = 100$, however, we lay off $ST - EI/n = 7.11$ inches of paper as a pole distance, and with a trial pole $O''$ in A diagram showing a continuous girder with a moment area diagram. 490 MECHANICS OF ENGINEERING. the vertical $TW$ draw the trial equilibrium polygon or elastic curve $A'B'$, and with it determine $\alpha$, then the final polygon $AB'$ as already prescribed. In $AB'$ we find the greatest ordinate, $NK$, to measure 0.88 inches of paper, which represents an actual distance of $0.88 \times 20 = 17.6$ inches. But the vertical dimensions of the exaggerated elastic curve $A'B'$ are so 100 times as great as those of the actual, hence the maximum deflection is $d=3.5 \times 0.176 = 0.6176$ in. [This maximum deflection could also be obtained from the oblique polygon $A"B"$ whose vertical dimensions are equal to those of $A'B'$. By the formula of § 235 we obtain $d=0.174$ inches.] 395. Direction of End-Tangents of Elastic Curve in the Forging Problem—As an illustration bearing on subsequent work let us suppose that the only result required in § 394 is $\tan a_0$, i.e., the tangent at the angle $B'TT'$, which the tangent-line $BT'$ to the elastic curve at the extremity $A'$, Fig. 343, makes with the horizontal line $AB'$, ($\tan a_0$ is called the "slope," at $A$). Let $FS'$ be the tangent-line at $E$. These two "end-tangents" are parallel respectively to $EF$ and $FO'$, and intersect at some point $R$. Now since $AKD$ is an equilibrium polygon sustaining an imaginary set of forces applied by the moment-area $ACB$, the axis of the moment-area $ACB$, the intersection $R$ must lie in the vertical containing the centre of gravity, $U$, of that moment-area [§ 336]. Hence, if the vertical containing $U$ is known in advance, or, as in the present case, is easily constructed without making the strip-subdivision of § 394, we may determine the end-tangents very briefly by considering the whole moment-area, $M.A.$ (considered as a load) applied in the vertical direction. Since $ACB$ is a triangle, we find $U$ by bisecting $AB$ in $X$, joining $CX$, and making $XU = \frac{1}{2}XC$, and then drawing a vertical through $U$. Laying off EF=6.4 inches (so as to represent a moment-area of 192,000 (sq. in.) (lbs.) on a scale of 30,000 (sq. in.) (lbs.) per linear inch of paper), CONTINUOUS GIRDER BY GRAPHICS. 491 and making $ST$=7.11 inches as before, we assume a trial pole $O'$ on $TF$, draw the two rays $OE'$ and $OF'$, construct the corresponding trial polygon by two segments $A'B'$, for the point $B'$, and $C'D'$, for the point $D'$. With the pole $O'$ on $TF$ and on a level with $n'$ we draw the two rays $OE'$ and $OF'$, and the corresponding segments $A'R$, and $RB'$. ($B'$ should be on a level with $A'$, as a check.) These two segments are the end-tangents required. We have, therefore, $$\tan a_0 = B'T' + A'T' = B'P' + A'P'$$ In the present numerical problem we find $B'T'$ to measure 3 in. of paper, i.e., 60 in. of actual distance for the exag. elastic curve, and therefore 0.60 in. in the real elastic curve (with $n$ = 100) $$\therefore \tan a_0 = \frac{0.60}{130} = 0.005$$ It is now evident that the position and direction of the end-tangents of the elastic curve lying between any two supports are independent of the mode of distribution of the moment-area so long as the amount of that moment-area and the position of its centre of gravity remains unchanged. This relation is to be of great service. 398. Re-arrangement of the Moment-Area.—As another illustration conducing to clearness in later constructions, let us determine by still another method the end-tangents of the beam of §§ 394 and 395. See Fig. 444. As already seen, their location is independent of the arrangement of the moment-area between. Let us re-arrange this moment-area, viz., the triangle $AOB$, in the following manner: By drawing $AX$ parallel to $BC$, and prolonging $BO$ to $p$ in the vertical through $A$, we may consider the original moment-area $ACB$ to be compounded of the positive mom.-area $VBA$, a parallelogram, with its gravity-vertical passing through $D$, the middle of the span ; of the negative mom.-area $VCA$, a triangle whose gravity-vertical passes 492 MECHANICS OF ENGINEERING. through $D_1$ making $AD_1 = \frac{1}{2} AD$; and of another negative mom.-area, the triangle $ABX$, whose gravity-vertical passes through $D_3$ at one-third the span from $B$. That is, the (ideal) positive load $ACB$ is the resultant of the positive load $(M.A.)_b$ and the two negative loads (or upward pulls) $(M.L.)_a$ and $(M.L.)_c$ and may therefore be replaced by them without affecting the location of the end-tangents, at $\varphi$ A diagram showing the forces acting on a beam with a negative moment area. A diagram showing the forces acting on a beam with a positive moment area. and $B.$ of the elastic curve $AB$. These three moment-areas are represented by arrows, properly directed, in the figure, but must not be confused with the actual loads on the beam (of which, here, there is but one, viz., $P$). From the given shapes and dimensions, since $ACB = 192,000$ (sq. in.) (lbs.), we easily derive by geometrical principles : \[ (M.A.)_b = +576,000 \text{ (sq. in.) (lbs.)} \] \[ (M.A.)_c = -96,000 \text{ " } \] \[ (M.A.)_a = -288,000 \text{ " } \] Hence, with a pole distance $EI + n = 7.11$ in. as before, and a “moment-load-line” formed of $1'2'' = (M.A.)_b$, (on scale of 30,000 (sq. in. lbs.) to one inch) $2'3'' = (M.A.)_c$, and $3'4'' = (M.A.)_a$ first with a trial pole $O''$, construct the trial A diagram showing the construction of a trial pole for determining the moment areas. CONTINUOUS GIRDER BY GRAPHics. 493 polygon $A''B''$, and find $\pi$ in usual way (§ 337); then take a pole $O'$ on the horizontal through $x'$ and the vertical $TW$, and draw the new polygon $A''123 B''$. It should pass through $B'$ on a level with $A'$ and $A'1$ and $B'3$ are the required end-tangents (of the exagg. elastic curve). [Note—If $B'$ were not at the same level as $A'$, but (say) 0.40 in. below it, $B'$ should be placed at $m$, a distance 0.4 x 100 = 2 inches (on the paper) below its present position, (since the distance scale is 130 and $n = 100$, in this case) and the "abutment-line" of final polygon would be $A'm$]. Of course, this special re-arrangement of the moment-area is quite superfluous in the present problem of a discontinuous girder (not built in), but considerations of this kind will be found indispensable with the successive spans of a continuous girder. 397. Positive and Negative Moment-Areas in Each Span of a Continuous Girder (Pristmatic).—In the foregoing problem of a discontinuous girder (covering one span only) not built in at the ends (otherwise it would be classed among continuous girders), we have to find the moment-curves for polygons of arbitrary $H$, is easily found by § 389 without use of the elastic curve, and the end moments are both zero; i.e., the moment-curve meets the beam in the end-verticals) but in each span of a continuous girder the end-moments are not zero (necessarily), and the points in the end-verticals where the moment-curve must terminate (for an assumed $H$) cannot be found without the use of the elastic curve (or of some of its tangents) of the whole beam, dependent, as it is, upon the loading on all the spans, and the heights of the supports. Let Fig. 445 show, in general, any one span of a prismatic continuous girder (prismatic; hence $I$ is constant), between two consecutive supports $A_0$ and $B_0$, $P_0$, $P_n$, etc., are the loads on the span. [If the displacement of $A_0$ relatively to the end-tangent at $B_0$, and the angle between the end-tangents (of elastic 494 **MECHANICS OF ENGINEERING.** curve) were known, the moment-curve or equilibrium polygon $FWC$ ($AB$ being the axis of beam) might be found by a process similar to that in § 390, but the elastic curves in successive spans are so inter-dependent that the above elements can not be found directly.) A diagram showing the moment-curves for a beam with supports at A and B. Fig. 68. We now suppose, for the sake of discussion, that the whole girder has been investigated (by a process to be presented) for the given loads, spans, positions of supports, etc., and then the moment-curves $FEW_1L$ found, with the corresponding stresses $M_1A_1$, $M_2A_2$, etc., in the figure and section area $H$. The horizontal line $AB$ represents the axis of the beam (for this purpose considered straight and horizontal) as an axis from which to measure the moment ordinates. Thus, the moment (of the stress-couple) at $A$ is = $H \times AP$; at $B$, $H \times BG$; at $E$ and $L$, zero (points of inflection). Now, according to the usual conceptions of analytical geometry, we may consider the portion $EW_1L$ of the moment-area, above $AB$ as positive, and those below, $AEF$ and $LBG$, as negative; but since not one of these three --- CONTINUOUS GIRDER BY GRAPHICS. 495 areas, nor the position of its gravity-vertical, is known in advance, since they are not independent of the other spans, a more advantageous re-arrangement of the moment-area may be made thus: Join $FG$ and $FB$, and we may consider the original moment-area replaced by the following three component areas: the positive moment-area $FEWGF$ (shaded by vertical lines); the negative triangular moment-area $ABF$; and the negative triangular moment-area $BFG$ (the negative moment-area being shaded by horizontal lines). (In subsequent paragraphs, the remaining two negative moment-areas will be implied those just mentioned.) These three moment-areas, treated as loads, each applied in its own gravity-vertical, and considered in any order, may be used instead of the real distributed moment-areas, as far as determining, or dealing with, the end-tangents of the elastic-curve at $A_1$ and $B_1$ concerned ($\S 390$), and the following advantages will have been gained: (1.) The amount of the positive moment-areas, $(M.A.)$, in Fig. 445, are determined by the triangle between the polygon $FEWGF$ and the shunt-line $F$ of the latter), and the position of its gravity-vertical, are independent of other spans, and can be easily found in advance, since this moment-area and gravity-vertical are the same as if the part of the beam covering this span were discontinuous and simply rest on the supports $A_1$ and $B_1$, as in $\S 389$. (2.) The gravity-vertical of the left-hand negative moment-area $(M.A.)$, is one-third the span from the left-end-vertical $A_1A'$; that of the right $(M.A.)$, an equal distance from the right-end-vertical $B_1B'$. (3.) The two (right and left) negative moment-areas are triangular, each having the whole span $l$ for its altitude, and for its base the intercept $AF$ (or $BG$) on which the end-moment depends. Hence, if the amounts of these negative moment-areas have been found in any span, we may compute the values which $AF$ and $BG$ must have for a given $H_1$, and thus determine the value of $H_2$ of the moment-curve of that span (for that value of $H_2$). 496 MECHANICS OF ENGINEERING. For example, if $(M.A.)$ has been found (by a process not yet given) to be $160,000$ (sq. in.) (lbs.) while $AB = l = 160$ in., then the moment $M$, which $AF$ represents, is computed from the relation $$\frac{(M.A.)}{2} \times AB \times M_A$$ or, $M_A = \frac{2 \times 160,000}{160} = 2000$. in. lbs. If $H$ has been chosen = 100 lbs. we put $H \times AF = 2000$ and obtain $AF = 20$ inches of actual distance, so that with a scale of 1/20 for distances $AF$ would be one linear inch of paper. (Of course, in computing $BG$ the same value of $H$ must be used.) With $H = 100$ lbs., then, and $F$ and $G$ as known points of the equilibrium polygon $FEW G$, it is easily drawn by the principles of § 941. We see that the two negative moment-areas are the only elements affected by the continuity of the girder, in this re-arrangement of the actual moment-areas. In the lower part of Fig. 445 $A'$ and $B'$ represent the extremities of the (exagg.) elastic curve, the vertical distance $A'B''$, of $B'$ from the horizontal through $A'$ (in case the two supports $A_1$ and $B_1$ are not at same level, as we have supposed for illustration) being laid off in accordance with the principle of § 148. Note.-It is now evident that if the "false polygon" (as it will be called) $A'123 B'$ has been obtained (and means for doing this will be given later) in which the first and last segments are the end tangents of the (exagg.)elastic curve, and which bears the same relation to the three moment areas just mentioned, as that illustrated in Fig. 444, we may proceed further to determine the amounts of $(M.A.)_1$ and $(M.A.)_2$ as follows, by determining the moment-area diagram: Having laid off the known $(M.A.)$ (or positive moment-area) = 23%, and $ST = EF + n$, a line parallel to 12 drawn through 3, determines the pole $O'$, through which paral... A diagram illustrating mechanics of engineering calculations. CONTINUOUS GIRDER BY GRAPHICS. 497 lets to $A1$ and $P3$ will fix $1'$ and $4'$ on the vertical $SV$, and thus determine $1'P-(M.A.)$ and $4'P-(M.A.)$. Their numerical values are then computed in accordance with the scale of the moment-area diagram. The polygon $\overline{A'B'C'D'}$ will be called the "false polygon" of the span in question, its end-segments being the end-tangents of the elastic curve. 398. Values of the Positive Moment-Area in Special Cases. For several special cases these are easily computed, and as an illustration, Fig. 446 shows a continuous girder, $AF$ A diagram showing a continuous girder with five spans labeled I, II, III, IV, V. The diagram includes points A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. Fig. 446 of five spans, all six supports on a level, and the weight of the beam neglected. At the extremities $A$ and $F$, as at the other supports, the beam is not built in, but simply touches each support at one point only. Therefore at $A$ and $F$ are zero the moment-curve must pass through $A$ and $F$, so that in the first span the left negative moment-area, and in the fifth span the right negative moment-area, are zero. The positive moment-areas are shaded. On the first span is placed a uniformly distributed load $W$ over the whole span $l$. : : the positive moment-area for that span is the same as in the case of Fig. 235 (see (1) § 297) and is shown represented by a parabolic segment whose area is two-thirds that of the circumscribing rectangle, its value is 498 MECHANICS OF ENGINEERING. $$\left(M.A.\right)_{y}=\frac{1}{2}\cdot\frac{1}{4}\cdot Wl\times l=\frac{1}{4}\cdot Wl^2 \quad . \quad (1)$$ (see eq. (2) § 342), while its gravity-vertical bisects the span. The only load on the second span is a concentrated one, $P'$, at distances $l_1$ and $l_2$ from the extremities of the span; hence the positive moment-area is triangular and has a value $$\left(M.A.\right)_{x'}=\frac{1}{2}\cdot P'\cdot l_1\cdot l_2 \quad . \quad (2)$$ as in § 393. Its gravity-moment may easily be constructed as in Fig. 443 (see § 397). The third span carries no load; hence its positive moment-area is zero, and the actual moment-area is composed solely of the two triangular negative moment-areas $CDH$ and $DHL$, the moment-curve consisting of the single straight line $HI$. The fourth span carries a uniform load $W^{IV}=w^{IV}l$, and has a positive moment-area $$\left(M.A.\right)_{y}=\frac{1}{4}\cdot W^{IV}(w^{IV}) \quad . \quad (3)$$ as in eq. (1), acting through the middle of the span (gravity-vertical). Since the fifth and last span carries no load, its positive moment-area is zero, the moment curve being the straight line $JF$, so that the actual moment-area is composed of the left-hand negative moment-area. As $F$ represents a reaction, section or pressure of the support must be from above downwards to prevent the beam from leaving the point $F$; i.e., the beam must be "latched down," and the reaction is negative. If the beams were built in at $A$ (or $F$) the moment at that section would not be zero, hence the left (or right) neg. moment-area would not be zero in that span, as in our present figure. But in such a case the tangent of the elastic curve would have a known direction at $A$ (or $F$) and the problem would still be determinate as will be seen. CONTINUOUS GIRDER BY GRAPHICS. 499 $A\ldots F$ gives an approximate idea (exaggerated) of the form of the elastic curve of the entire girder. A change in the loading on any span would affect the form of this curve throughout its whole length as well as all of the moment curves. NOTE.-It is important to remark that any two of the triangular negative moment-areas which have a common base (hence lying in adjacent spans) are proportional to their altitudes, i.e., to the lengths of their verticals; they occur thus the same areas (Fig. 440) $GCH$ and $DCH$ have a common base $CH$, $\therefore \frac{(M.A.)_{CH}^{m}}{(M.A.)_{CH}^{m'}}=\frac{P}{l_{m}}\quad\ldots\quad(4)$ (The notation explains itself ; see figure). It also follows, that the resultant of these two neg. moment-areas (if required in any construction ; see § 400) acts in a vertical which divides the horizontal distance between their gravity verticals in such a ratio as that of the spans to which they belong [§ 21 and eq. (4) above]. Hence, since this horizontal distance is $\frac{1}{3}l^{*}+\frac{1}{3}l^{*'}$ their resultant must act in a vertical $Y^{**}$ whose distance from the gravity-vertical of $GCH$ is $\frac{1}{3}l^{*'}$, and from that of $DCH$, $\frac{1}{3}l^{*}$. 399. Amount and Gravity-Vertical of the Positive Moment Area of One Span as Due to Any Loading.--Since we can not deal directly with a continuous load by graphics, but must subdivide it into a number of detached loads sufficiently numerous to give a close approximation, let us suppose that this has already been done if necessary, and that $P_1$, $P_2$, etc., are the detached loads resting in the span $AB$ in question ; see Fig. 447. Since (by (1), § 397) the positive moment-area is the same as the total moment-area would be if this portion of the beam simply rested on the extremities of the span, not extending beyond them, then we have the convenient rule : 389 for finding it, remembering that in that paragraph the 500 MECHANICS OF ENGINEERING. oblique polygon in the lower part of Fig. 439 will serve as well as the (upper) one whose abutment-line is in the beam itself, as the moments are concurrent. Hence, Fig. 447, lay off the line $LL'$, take any pole $O$, with any convenient pole-distance $H$, and draw the equilibrium polygon $FWG$. After joining $FG$, $FWGF$ will be the positive moment-area required. To find its gravity vertical, divide the span $AB$, or $FG'$, into ten to twenty equal parts (each = $de$) and draw a A diagram showing a polygonal structure with various lines and points labeled. vertical through the middle of each. The lengths $e_1$, $e_2$ etc., on these verticals, intercepted in the moment-area, are proportional to the corresponding strips of moment-areas, each width being $\frac{1}{10}$ of an inch. Form a loose line $SK$, of the successive $e_i$'s, and with any pole $O'$, draw the equilibrium polygon $A'B'$ (for the z-verticals). The intersection, $R$, of the extreme segments, is a point in the required gravity-vertical ($\S$ 336). The amount of the moment-area is $(M.A.) = \Sigma(H.de)$ $$= H.de = H.de(e_1 + e_2 + e_3 + \ldots)$$ CONTINUOUS GIRDER BY GRAPHICS. 501 For example, with the span $l = 120$ inches, subdivided into twelve equal $d$, we have $d = 10$ inches (of actual distance). If $H = 4$ inches of paper and $SK = 10.2$ inches of paper, the force-scale being 80 lbs. to the inch, and the distance-scale 15 inches to the inch ($1:15$), we have $$M.A._{h} = [4 \times 80] \times 10 \times [10.2 \times 15] = 489600. \text{(sq. in.)}$$ 400. Construction of the "False-Polygon" For All the Spans of a Given (Pristmatic) Continuous Girder, Under Given Loading, and With Given Heights of Supports.—[See note in § 397 for meaning of "false-polygon".]—[See note in § 397 for meaning of "false-polygon".]—[See note in § 397 for meaning of "false-polygon".] We suppose that the given girder covers three unequal spans, Fig. 448, with supports at unequal heights, and that both extremities $A$ and $D$ are built-in, or "fixed", horizontally. To clear the ground for the present construction, we suppose that, from the given loading in all the spans, the effective lengths of each span have been obtained in numerical form (so many (sq. in.) (lbs.) or (sq. in.) (tons)) and its gravity-vertical determined by § 398 or § 399; that the horizontal distances (i.e., the spans $t$, $r$, and $l'$ and the distance between the above gravity-verticals and the supports) have been laid off on some convenient scale; that $EI$ has been computed from the material and shape of section of the girder and expressed in the same units as the above moment-areas; that the verticals through the points selected (since $EI \cdot n$ is to be the pole distance of all the moment-area-diagrams), and that the vertical distances of $B$, $C$, and $D$, from the horizontal through $A$, have been laid off accordingly (see note in § 396). In the figure (448) verticals are drawn through the points of support; also verticals dividing each span into thirds, since the unknown negative moment-areas (subscripts 1 and 3) set in the latter ($§ 397$); and verticals of the known positive moment-areas. The verticals $T_1V_1$, $T_2V_2$, etc., will be discussed later. The problem may now be stated as follows: Given the positions of the supports, the value of $EI$ and $\alpha$. 502 MECHANICS OF ENGINEERING. the fact that the girder is fixed horizontally at $A$ and $D$, the heights of supports, the location of the gravity-vertices of all the positive and negative moment-areas, and the amounts of the positive moment-areas ; it is required to find graphically the "false-polygon" in each span. The "false-polygons," viz.: $A123B$ for the first span (on the left), $D123C$ for the second, etc., are drawn in the figure A diagram showing three false polygons labeled (M.A.)' for spans A123B, D123C, and E123F. for the purpose of discussing their properties at the outset. Since $3B$ and $B1$ are both tangent to the elastic curve at $B$, they form a single straight line ; similarly $C1$ is but the prolongation of $3C$. Also $A1$ and $D1$ must be horizontal since the beam is built in horizontally at its extremities $A$ and $D$. That is, the three false polygons form a continuous equilibrium polygon $A...D$, in equilibrium under the "loads" $(M.A.)'$, $(M.A.)'$, $(M.A.)'$, etc., so that we might use a single mom.-area-diagram in connection with it, but for convenience the latter will be CONTINUOUS GIRDER BY GRAPHICS. 503 drawn in portions, one under each span, with a pole distance $= \frac{5f}{2}$. Of this polygon $A\ldots D$, we have the two segments $A1$ and $B1$ already drawn, and know that it passes through the points $B$ and $C$; we shall next determine by construction other points (called "fixed points") $p_1^{'}, p_2^{''}, p_3^{'''},$ and $p_4^{''''}$ (in the prolongations of) certain other segments. To find the "fixed point" $p_1^{'},$ where the segment $23$ in the first span cuts the vertical $V'$, the gravity-vertical of $(M.A.)'$. The vertex $p_1^{'},$ or 1, is already known, being the intersection of $A1$ with $V'$. Lay off $23 = (M.A.)'$ which is known, on the line of action of the resultant of $O_1^{'}, O_2^{'}, O_3^{'}, O_4^{'''}$, join $O_1^{'}, O_2^{'},$ and $O_3^{'},$ 12 to 20 to 20', to determine $\bar{a}_2$, on the vertical $(M.A.)'$, then through $\bar{a}_2$ line to $O_4^{'''}$ to cut $V'$ in $p_4^{''''}$. The unknown segment 23 must cut $F'$ in the same point ; since all positions of $O_4^{'''}$ on the vertical $T'U'$ will result in placing $p_4^{''''}$ in this same point, and one of these positions must be the real pole $O'$ (unknown). [This is easily proved in detail by two pairs of similar triangles]. To determine the "fixed-point" $p_2^{''}$, in the prolongation of segment 12 of second span. The prolongations of the segments 12 (of first span) and 12 (of second span) must meet in a point $\bar{k}$ in the (vertical) line of action of the resultant of $(M.A.)'$ and $(M.A.)'$ (§ 336). Although the amounts of $(M.A.)'$ and $(M.A.)'$ are unknown, still the vertical line of action of their resultant (by § 398, Note) is known to be $Y'$, a horizontal distance $\frac{5f}{2}$ to the right of $(M.A.)'$; hence $Y'$ is easily drawn. Therefore, the unknown triangle $\bar{k}31$ has its three vertices on three known verticales, the side $\bar{k}3$ passes through the known point $p_3^{'},$ and the side 13 through the known point $B$. Now by prolonging $\bar{a}_2$, $\bar{a}_3$, (or any line through $p_3^{'})$ to cut $(M.A.)'$ and $Y'$ in $\bar{a}_2$ and $\bar{a}_3$, respectively, joining $\bar{b}_3$ and prolonging this line to cut $(M.A.)'$ in some point $\bar{l}_o$ and then joining $\bar{e}_o$, we have a triangle $\bar{k}\bar{e}_o\bar{l}_o$ of which we can make a 504 MECHANICS OF ENGINEERING. statement precisely the same as that just made for $k'31$. But if two triangles [as $k'31$ and $k_3'3_1$] have their vertices on three parallel lines (or on three lines which meet in a point) the three intersections of their correspond- ing sides must lie on the same straight line [see reference to Chaveneut, § 378 a)]. Of these intersections we have two, $p_3'$ and $B$, hence the third must lie at the intersection of the line $p_3'B$ (prolonged) with $k_3'3_1$, and in this way the "fixed point" $p''$, which was found in the segment 12 (of second span) prolonged, is found. Draw a vertical through it and call it $V''$. The fixed point $p_3''$ (in prolongation of segment 25 of sec- ond span) lies in the vertical $V''$ and is found from $p'$ and the known value of $(M.A.)_{\mathrm{v}}''$ precisely as $p_3''$ was found from $p'$. That is, we lay off vertically $2^{\circ} 5^{\prime\prime}$ precisely as $(M.A.)_{\mathrm{v}}$, and join 2° and 5° to $O_1$, which is any point at distance $EI = a$ to the right of $2^{\circ}$. Through $p'$ draw a line to $2^{\circ} O_1$ to cut (at its foot) then $2^{\circ} 1$ to $O_1$ 5° to determine $p''$ on the vertical $V''$. The fixed points $p'''$ and $p_3'''$ in the third span lie in the prolongations of the segments 12 and 23, respectively, of that span, $p'''$ being found from the points $p_3''$ and $C$ and the verticals $(M.A.)_{\mathrm{v}}'', Y''$, and $(M.A.)_{\mathrm{v}}'''$, in the same manner as $p'''$ was determined with similar data, while $p_3'''$, in the same vertical $V'''$ as $p'''$, depends on $(M.A.)_{\mathrm{v}}'''$ and its gravity vertex as already illustrated; hence the detail need not be drawn in figure. In this way for any number of spans we proceed from span to span toward the right and determine the succes- sive fixed points, until the points $p$ and $p_3$ of the last span have been constructed, which are $p''$ and $p_3''$ in our present problem. Since $p_3'''$ is a point in the segment 23 (prolonged) of the last span, we have only to join it with 3 in that span, a point already known, and the segment 23 is determined. Joining the intersection 2 with $p'''$ we determine the new point 21 of course by construction as before, which is then joined with C and prolonged to intersect $(M.A.)_{\mathrm{v}}'$ to fix the segment IC3 and the point 3. Join CONTINUOUS GIRDER BY GRAPHICS. 505 3 $p_{n}^{\prime}$, and proceed in a similar manner toward the left until the whole equilibrium polygon (or series of "false polygons") is finally constructed; the last step being the joining of 2 with $p'$. **401. Treatment of Special Features of the Last Problem.—(L)** If the beam is simply supported at A, Fig. 448, instead of built-in, $(M.A.)_t'$ becomes zero, and the two segments $A1$ and 12 of that span form a single segment of unknown direction. Hence, the point $A$ will take the place of $p'$, and the vertical $F_a$ that of $P'$. (2.) Similarly, if $D$, in the last span, is a simple support (beam not built in) $(M.A.)_t'$ becomes zero, and the segments 13 and 23 will form a single segment of unknown direction, so that after $p'$ has been found, we join $p''$ and $D$ to determine the segment $D2$ ; i.e., in this last span, $D$ takes the place of 3 of the previous article. (3.) If the first span carries no load $(M.A.)_t'$ is zero, and the segments 12 and 23 will form a single segment 23. Hence if the beam is built in at $A$, $p_t$ will coincide with the known point $p'$ (i.e., 1), while if $A$ is a simple support $p$ and $p'$ coincide with $A$, since then $(M.A.)_t'$ is zero and $\overline{123}$ is a straight line. (4.) If the last span is unloaded (third span in Fig. 448), $(M.A.)_t''$ is zero, 123 becomes a single segment, and hence $p''$ will coincide with $p'''$; so that after $p'''$ has been constructed it is to be directly joined to $3$, if the beam is built in at $D$, and will thus determine the segment 13 ; or to $D$, if $D$ is a simple support, (for then $(M.A.)_t''$ is zero and the three segments 12, 23, and 3 form a single segment.) (5.) If an intermediate span is unloaded (say the second span, Fig. 448) the positive mom-area, $(M.A.)_t''$, is zero, 123 becomes a straight line, i.e. a single segment, and therefore $p_n''$ coincides with $p'''$; hence, when $p'''$ has been found we proceed as if it were $p''$". **402. To Find the Negative Mom.-Areas, the Mom.-Curves, Shears, and Reactions of the Supports.—(L)** Having constructed the 306 MECHANICS OF ENGINEERING. false polygons according to the last two articles, the negative moment-areas of each span are then to be found by the note in § 397, Fig. 445, and expressed in numerical form. [If the positive mom.-area of the span is zero the points $2^{\prime}$ and $3^{\prime}$ will coincide, Fig. 445, and in the case mentioned in (3), (or (4)), of § 401, if $A$ (or $D$) were a simple support, in Fig. 445, the mom.-area-diagram of Fig. 445 would have but two points.] (2.) The moments at the supports (or "end-moments" of the respective spans) depending, as they do, directly on the negative mom.-areas, can now be computed as illustrated in (3) § 397. The fact that each "end-moment" may be obtained from two negative mom.-areas, separately, one in each adjacent span (except, of course at the extremities of the girder) forms a check on the accuracy of the work. The two values should agree within one or two per cent. (3.) The mom.-curve of each span is formed from a force-diagram whose load-line consists of the actual loads on the span laid off in proper order, can now be drawn, a convenient value for $H$ having been selected (the same $H$ for all the spans, that the moment-curves of successive spans may form a continuous line for the whole girder); since we may easily compute the proper moment ordinate at each support to represent the actual moment, then, for the $H$ adopted, by (3) § 397. The moment-curve of each span, since we know its two extreme points and its pole-distance $II$, is then constructed by § 341. (4.) The shear. Since the last construction involves drawing the special force-diagram for each span, with a ray corresponding to each part of the span between two consecutive loads, the shear at any section of the beam is easily found as being the length of the vertical projection of the "proper ray," interpreted by the force-scale of the force-diagram, as in §§ 389 and 390. With the shears as CONTINUOUS GIRDER BY GRAPHICS. 507 ordinates a shear-diagram may now be constructed, if desired, for each span. The directions of the shears should be carefully noted. (5.) Reactions of supports. Let us consider "free" the small portion of the girder, at each point of support, between two sections, one close to the support on each side, Fig. 449. Suppose it is the support $C$, and call the reaction, or pressure at that support $R_c$. Then, for vertical equilibrium, i.e., $Y = 0$ (§ 36), we have $$R_c = J_{cn} + J_{cn} \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \\ R_c = J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cn} - J_{cn} + J_{cm}-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_cm-J_cm+J_km-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1cm-1 cm 508 MECHANICS OF ENGINEERING. spans of 16, 20, and 14 feet respectively, and bearing a single load in each of the extreme spans, but a uniform load over the entire central span. As indicated thus :
7 ft. 9 ft. 20 ft. 8 ft. 6 ft.
A B C D
P' P" P'"
30 tons 40 tons 32 tons
(This is a practical case where $W''$ is the weight of a brick wall, and $P'$ and $P'''$ are loads transmitted by columns from the upper floors of the building; $A$ and $D$ are simple supports, and the weight of the girder is neglected.) The beam has a moment of inertia $I = 707$ square inches, and the modulus of elasticity of the iron is $E = 25,000$ lbs per sq. in., = 25,000 tons per sq. in. Although the graphical statics of this simple girder under given loading, with supports at the same level, it may easily be shown that the moments, shears, and reactions, to be obtained graphically, are the same for all values of $I$, so long as the elastic limit is not surpassed, still, on account of the necessity of its use in other problems (supports not on a level), we shall proceed as if the value of $I$ were essential in this one. Selecting the inch and ton as units for numerical work, we have $$EI = 12500 \times 707 = 8,887,500 \text{ (sq. in.) (tons)}$$ while the respective positive mom.-areas, from eqa. (2) and (3) of § 309, are: $$\begin{align*} (M.A.)' &= \frac{1}{2} \times 30 \times 84 \times 108 = 136,080 \text{ (sq. in.) (tons)} \\ (M.A.)'' &= \frac{1}{4} \times 40 \times 240^{\circ} = 199,000 \text{ " " " } \\ (M.A.)''' &= \frac{1}{2} \times 32 \times 96 \times 72 = 110,392 \text{ " " " } \end{align*}$$ Adopting a scale of 60,000 (sq. in.)/(tons) to the linear inch of paper, for mom.-area diagrams, we have for the above mom.-areas 2.27 in., 3.20 in., and 1.84 in., respectively, on the paper, for use in Fig. 448. CONTINUOUS GIRDER BY GRAPHICS. 509 Having laid out the three spans on a scale of 60 inches to the inch of paper, with $A$, $B$, $C$, and $D$ in the same horizontal line, we find by the construction of Fig. 443, that the gravity-vertical of $(M.A.)_y'$ lies 3.6 in. to the left of the middle in the first span, that of $(M.A.)_y''$ is in the right third point of the third span; while that of $(M.A.)_y'''$, of course, bisects the central span. Hence we draw these verticals; and also those of the unknown negative mom.-areas through the one-third points; renumbering (§ 401, (1) and (2)) that $(M.A.)_y'$ and $(M.A.)_y''$ are both zero in this case. Since $EI = 8,937,500$ (sq. in.) (tons), it would require 147.29 in. to represent it, as pole-distance, on a scale of 60,000 (sq. in.) (tons) to the inch; hence let us take $e = 50$ for the degree of (vertical) exaggeration of the false polygons, since the corresponding pole-distance $\frac{EI}{e^2} = 2.04$ in. of paper is a convenient length for use with the values of $(M.A.)_y'$, $(M.A.)_y''$, etc., above given. Following the construction of Fig. 448, except that $p'$ is at $A$, and $p''$ is to be joined to $D$ (§ 401), (the student will do well to draft the problem for himself, using the prescribed scales,) and thus determining the false-polygons, we then construct and compute the sag. mom.-areas according to § 402 (1), and the note in § 397, obtaining the following results: \begin{align*} (M.A.)_y', \text{1.43 in. of pap.,} &= 85,800 \text{ (sq. in.) (tons)} \\ (M.A.)_y'', \text{1.77 " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " } &= 106,200 \text{ " " } \\ (M.A.)_y''', \text{1.68 " " " " " " " " " " " " " " " " " " } &= 100,800 \text{ " } \\ (M.A.)_y''''', \text{1.17" " " " } &= 70,200 \text{ " } \end{align*} The remaining results are best indicated by the aid of Fig. 450. Following the items of § 402, we find ((§ 397) that the moment at $B$, using $(M.A.)_y''$, is $$M_B = \frac{5 \times 106,200}{240} = 885 \text{ inch-tons}.$$ 510 MECHANICS OF ENGINEERING. [or, using $(M.A.)_x$, $M_a = \frac{2 \times 85,000}{192} = 893$ in. tons.] Similarly, $M_c = \frac{2 \times 100,800}{240} = 840$ in. tons. [or, using $(M.A.)_y$", $M_c = 835.7$ in. tons.] Hence, taking means, we have, finally, $$M_a = 0; M_b = 889 \text{ in. tons}; M_c = 837.8; M_d = 0.$$ Fig. 450 shows the actual mom.-areas and shear-diagrams, which are now to be constructed. A diagram showing the moment areas and shear diagrams for a beam with varying cross-sectional areas. Selecting a value $H = 20$ tons for the pole-distance of the successive force-diagrams (the scale of distances being 5 ft. (60 in.) to the inch we have [(3) § 397]) $$20 \times BG = M_a = 889 \text{ in.-tons} ; BG = 44.4 \text{ in. of actual distance}, \text{ or } 0.74 \text{ in. of paper; also } 20 \times CK = M_c = 837.8 \text{ in.-tons} ; CK = 41.89 \text{ in., or } 0.698 \text{ in. of paper}.$$ Having thus found $G$ and $K$, and divided $BC$ into ten equal parts, applying four tons in the middle of each, we construct by § 341 an equilibrium polygon which shall pass through $G$ and $K$ and have 20 tons as a pole-distance (We take a force-scale of 10 tons to the inch.) It will form a succession of short tangents to a parabola, and is the moment curve for span $BC$. Similarly, for the single CONTINUOUS GIRDER BY GRAPHics. Load $P'$ and $P''$ in the other two spans, we draw the equilibrium polygons $AN'G$ and $KZ'D$, for the same $H$ as before, and passing through $A$ and $G$, and $K$ and $D$, respectively. Scaling the moment-ordinates $NN'$, $QQ'$, and $ZZ''$, reducing to actual distance and multiplying by $H$, we have for these local moment maxima, $M_0 = 1008$, $M_1 = 386$, and $M_2 = 398$, in. tons. Evidently the greatest moment is $M_2$ and . . . the stress in the outer fibre at $\mathcal{N}$ will be ($\S 289$) $$p_n = \frac{M_{0e}}{f} = \frac{1008 \times 7\frac{1}{2}}{707} = 10.0 \text{ tons per sq. inch which is much too large.}}$$ If we employ a 20-inch heavy beam, with $I = 1650$ biquad. in., the preceding moments will still be the same (supports all at same level) and we have $$p_s = \frac{1008 \times 10}{1650} = 6.06 \text{ tons per sq. in.}},$$ or nearly 12,000 lbs. per sq. in., and is therefore safe ($\S 183$). If three discontinuous beams were to be used, the 20-inch size of beam (heavy) would be much too weak, in each of the three spans, as may be easily shown; hence the economy of the continuous girder in such a case is readily perceived, but it can be seen that the continuity and smoothness do not differ so much in the shear-diagrams as in the moment curves. By scaling the vertical projection of the proper rays in the special force diagrams (as in $\S\S 389$ and 390) we obtain the shear for any section on $AN$, as $J_{\mathcal{N}}$ (see Fig. 449 for notation) = 12.3 tons; on $NB$, $J_{\mathcal{N}} = 17.7$ tons; from $B$ to $C$ it varies uniformly from $J_{\mathcal{N}} = 20.3$ tons, through zero at $Q$, to $J_{\mathcal{L}} = 19.7$ tons of opposite sign. Also, for OZ, $J_{\mathcal{N}} = 15.8$ tons; for OC, $J_{\mathcal{L}} = 14.2$ tons. Hence, the reactions of the supports are as follows: $$R_A = J_{\mathcal{N}} - J_{\mathcal{L}} = 12.3 - (-19.7) = 32.0 \text{ tons},$$ $$R_B = J_{\mathcal{N}} + J_{\mathcal{L}} = 20.3 + (-15.8) = 4.5 \text{ tons},$$ $$R_C = J_{\mathcal{L}} - J_{\mathcal{N}} = -19.7 - (-14.2) = -5.5 \text{ tons}. $$ 512 MECHANICS OF ENGINEERING. [In the shear-diagram, the shear-ordinates are laid off below the axis when the shear points down, the "free-body" extending to the right of the section considered, (as $I_0$ in Fig. 449) and above, when the shear points upward for the same position of the free body.] If we divide the max. shear, 20.3 tons by the area of the web, 13.75 sq. in., of the 20-inch heavy beam, ($\S$356), we obtain 1.5 tons or 8000 lbs. per sq. in., which is < 4000 ($\S$ 185). Notice the points of inflection, $\nu$, $\psi$, etc., where $M$ is zero. Suitable bearing surface should be provided at the supports. A swing-bridge offers an interesting case of a continuous girder. 404. Continuous Girders of Variable Mom. of Inertia.--If $I$ is variable and $L_0$ denote the mom. of inertia of some convenient standard section, then we may write $I = I_0 + m$, when $m$ denotes the number of times $L_0$ contains $L$. In a non-prismatic beam, $m$ is different for different sections but is easily found, and will be considered given at each section. In eq. (1) of § 391, then, we must put $I_0 + m$ in place of $I$ and thus write $$\frac{d y}{dx^2} = \frac{[m M dx]}{h L} \qquad \qquad (1)$$ and (pursuing the same reasoning as there given) may therefore say that in a girder of variable section if each small vertical strip (Mdx) of the moment-area be multiplied by the value of $m$ proper to that section, and these products (or virtual mom.-area strips) considered as loads, the elastic curve is an equilibrium polygon for those loads with a pole distance = $E L_0$. In modifying § 400 for a girder of variable section, then, besides taking $E L_0 + m$ as pole distance, proceed as follows: Construct the positive mom.-area for each span according to § 399; for each $x$ of Fig. 447, substitute $m$ (each : CONTINUOUS GIRDER BY GRAPHics. 513 having in general its own $m$, and thus obtain the "virtual positive mom.-area," and its gravity vertical. Similarly, there will be an unknown "virtual neg. mom.-area," not triangular, replacing each neg. mom.-area of § 400. Though it is not triangular, each of its ordinates equals the corresponding ordinate of the known triangular neg. mom.-area multiplied by the proper $m$ and its gravity-vertical (which is independent of the amount of the unknown neg. mom.-area) is found in advance by the process of Fig. 447, using, for $s'$, a set of ordinates obtained thus: Draw any two straight lines $AB$ and $FB$, Fig. 445, (for a left-hand trial neg. mom.-area; or $FB$ and $GF$ for a right-hand one) meeting in the end-vertical of the span, divide the span into ten or twenty equal spaces, draw a vertical through each point of division, and join these to $AB$ and $FB$. Add those intercepts and call the sum $S$. Multiply each intercept by the proper $m$, and with these new values as $s'$ construct their gravity vertical as in Fig. 447. Add these new intercepts, call the sum $R_s$, and denote the quotient $S + R_s$ by $\beta$. We substitute the three verticals mentioned, therefore, for the mom.-area verticals of § 400, and the "virtual posi. mom.-area" for the pos. mom.-area, in each span; proceed in other words, we have substituted the polynomials according to § 400. Then the result of applying this construction in the note § 397 will be the "virtual neg. mom.-areas," each of which is to be multiplied by the proper $\beta$ to obtain the corresponding triangular neg. mom.-areas, with which we then proceed, without further modifications in the process, according to (2), (3), etc., of § 402. (The Theorem of these "virtual mom.-areas" is due to Prof. Edly, see p. 96 of his Researches in Graphical Statics,) and so far as I am aware of all works on this subject. 405 Remarks—It must be remembered that any unequal settling of the supports after the girder has been put in place, may cause considerable changes in the values of the moments, shears, etc., and thus cause the actual stresses to be quite different from those computed without taking 514 MECHANICS OF ENGINEERING. into account a possible change in the heights of the supports. See § 271. For example, if some of the supports are of masonry, while others are the upper extremities of high iron or steel columns, fluctuations of length in the latter due to changes of temperature will produce results of the nature indicated above. If an open-work truss of homogeneous design from end to end (treated as a girder of constant moment of inertia, whose value may be formulated as in § 388.) is used as a continuous girder under moving loads, it will be subject to "reversal of stress" in some of its upper and lower horizontal members, i.e., the latter must be of a proper design so that both their tension and compression (according to the position of the moving loads,) and this may disturb the assumption of homogeneity of design. Still, if $I$ is variable, § 404 can be used; but since the weight of the truss must be considered as part of the loading, several assumptions and approximations may be necessary before establishing satisfactory dimensions. PART IV. HYDRAULICS. CHAPTER I. DEFINITIONS--FLUID PRESSURE--HYDROSTATICS BEGUM. 400. A Perfect Fluid is a substance the particles of which are capable of moving upon each other with the greatest freedom, absolutely without friction, and are destitute of mutual attraction. In other words, the stress between any two contiguous portions of a perfect fluid is always one of compression and normal to the dividing surface at every point; i.e., no shear or tangential action can exist on any imaginary cutting plane. Hence if a perfect fluid is contained in a vessel of rigid material the pressure experienced by the walls of the vessel is normal to the surface of contact at all points. For the practical purposes of Engineering, water, alcohol, mercury, air, steam, and all fluids may be treated as perfect fluids within certain limits of temperature. 407. Liquids and Gases.--A fluid a definite mass of which occupies a definite volume at a given temperature, and is incapable both of expanding into a larger volume and of being compressed into a smaller volume at that temperature, is called a Liquid; of which water is a very common concrete example; whereas a Gas is a fluid a mass of which is capable of almost indefinite expansion or compression, according as the space within the confining vessel is made larger or smaller, and always tends to fill the vessel, which must therefore be closed in every direction to prevent its escape. 515 516 MECHANICS OF ENGINEERING. Liquids are sometimes called inelastic fluids, and gases elastic fluids. 408. Remarks.—Though practically we may treat all liquids as incompressible, experiment shows them to be compressible to a slight extent. Thus, a cubic inch of water under a pressure of 15 lbs on each of its six faces loses only fifty millionths (0.0000005) of an inch in volume when heated at the same temperature; if the temperature be sufficiently raised, however, its bulk will remain unchanged (provided the initial temperature is over 40° Fahr.). Conversely, by heating a liquid in a rigid vessel completely filled by it, a great bursting pressure may be produced. The slight cohesion existing between the particles of most liquids is too insignificant to be considered in the present connection. The property of indefinite expansion, on the part of gases, by which a limited mass of gas can continue to fill a confined space which is progressively enlarging, and exert pressure against its walls, is satisfactorily explained by the “Kinetic Theory of Gases,” according to which the gaseous particles are perfectly elastic and in continual motion, impinging against each other and the confining walls. Nevertheless, for practical purposes, we may consider a gas as a continuous substance. Although by the abstraction of heat, or the application of great pressure, or both, all known gases may be reduced to liquids (some being even solidified); and although by converse processes (imparting heat and diminishing the pressure) liquids may be transformed into gases, the range of temperature and pressure in all problems to be considered in this work is supposed kept within such limits that no extreme changes of state, of this character, take place. A gas approaching the point of liquefaction is called a Vapor. Between these two extreme states we find all grades of intermediate conditions of matter. For example, some substances are described as soft and plastic solids, as soft putty, moist earth, pitch, fresh mortar, etc.; and others as viscous and sluggish liquids, as molasses and glycerine. In sufficient bulk, A page from a book with text discussing mechanics of engineering. DEFINITIONS—FLUID PRESSURE—HYDROSTATICS. 517 however, the latter may still be considered as perfect fluids. Even water is slightly viscous. 409. Heaviness of Fluids.—The weight of a cubic unit of a homogeneous fluid will be called its heaviness, or rate of weight (see § 7), and is a measure of its density. Denoting by $y$, and the volume of a definite portion of the fluid by $V$, we have, for the weight of that portion, $$G = yV.$$ .......................... (1) This, like the great majority of equations used or derived in this work, is of homogeneous form (§ 6), i.e., admits of any system of units. E.g., in the metre-kilogram-second system, if $y$ is given in kilos per cubic metre, $V$ must be expressed in cubic metres, and $G$ will be obtained in kilos; and similarly in any other system. Thus, if $y$, and $V$, are respectively one dimension of force divided by the dimension of length. In the following table, in the case of gases, the temperature and pressure are mentioned at which they have the given heaviness, since under other conditions the heaviness would be different; in the case of liquids, however, for ordinary purposes the effect of a change of temperature may be neglected (within certain limits). HEAVINESS OF VARIOUS FLUIDS.* [In ft. lb. sec. system; $y =$ weight in lbs. of a cubic foot.]
Liquids Gases* (At. temps. of melting ice: and 14.7 lbs. per sq. m. tension.)
Peach water. $y = 63.5$ lbs.
Fee water. $y = 64.0$ lbs.
Mercury. $y = 848.7$ lbs.
Acohol. $y = 848.7$ lbs.
Crude Petroleum, about $y = 55.0$ lbs.
(N.P.) $y = 0.008624$ lbs., and a cubic foot 1000 av. on.)
Atmospheric Air. $y = 0.0873$ lbs.
Oxygen. $y = 0.0895$ lbs.
Nitrogen. $y = 0.0798$ lbs.
Helium. $y = 0.0092$ lbs.
Illuminating gas from $y = 0.0300$ lbs.
lamp. $y = 0.0495$ lbs.
Natural Gas, about $y = 0.0500$ lbs.
* See Trautwine's Civ. Engineer's Pocket Book for an extended table—p. 380, edition of 1885. 518 MECHANICS OF ENGINEERING. For use in problems where needed, values for the heaviness of pure fresh water are given in the following table (from Rossetti) for temperatures ranging from freezing to boiling; also the relative density, that at the temperature of maximum density, 39° 3 F. being taken as unity. The temperatures are Fahr., and y is lbs per cubic foot.
Temp. Rel. Density y Temp. Rel. Density y Temp. Rel. Density y
32° 0.9967 62.416 60° 0.9967 61.900 140° 0.9886 61.396
35° 0.9941 62.416 62° 0.9941 61.900 150° 0.9845 61.308
38° 1.0000 62.424 80° 0.9969 61.721 160° 0.9728 61.008
40° 0.9957 62.424 82° 0.9957 61.721 170° 0.9678 61.298
43° 0.9997 62.432 100° 0.9915 61.696 180° 0.9578 61.354
45° 0.9975 62.432 120° 0.9870 61.671 210°
From D. K. Clark's for temp = 18°, 25°, 32°, 45°, 55°, 65°, 75°, 85°, 95°, 105°, 115°, 125°, 135°, 145°, "Manual." \[ y = \frac{G}{V} = \frac{y}{\frac{G}{V}} = \frac{y}{\frac{G}{V}} \] EXAMPLE 1. What is the heaviness of a gas, 432 cub in. of which weigh 0.368 ounces? Use ft.-lb.-sec system. 432 cub in. = \(\frac{4}{4}\) cub ft.; and 0.368 oz = 0.028 lbs. \[ \therefore y = \frac{G}{V} = \frac{y}{\frac{G}{V}} = \frac{y}{\frac{G}{V}} \] EXAMPLE 2. Required the weight of a right prism of mercury of 1 sq. inch section and 30 inches altitude. \(V = 30 \times 1 = 30\) cub in.\(= \frac{30}{\frac{G}{V}} = \frac{G}{V}\) \(y\) for mercury is \(848.7\) lbs per cub ft. \[ \therefore \text{its weight} = G = VY = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] \[ = \frac{G}{V} = \frac{G}{V} \] 410 Definitions.—By Hydraulics we understand the mechanics of fluids as utilized in Engineering. It may be divided into Hydrostatics, treating of fluids at rest; and Hydrodynamics (or Hydrokinetics), which deals with fluids in motion. (The same Pneumatics is sometimes used to cover both the statics and dynamics of gaseous fluids.) DEFINITIONS—FLUID PRESSURE—HYDROSTATICS. 519 [Rankine's nomenclature has been adopted in the present work. Some recent writers use the term Hydromechanics for mechanics of fluids, subdividing it into Hydrotatics and Hydrokinetics, as above; they also use the term Dynamics to embrace both of the two divisions called Statics and Kinetics respectively. Though unusual, perhaps, the term Hydraulics is here used to cover the applied Mechanics of Gases as well as of Liquids.] Before treating separately of liquids and gases, a few paragraphs will be presented applicable to both kinds of fluids. 411. Pressure per Unit Area, or Intensity of Pressure.—As in § 180 in dealing with solids, so here with fluids we indicate the pressure per unit area between two contiguous portions of fluid, or between a fluid and the wall of the containing vessel, by $p$, so that if $dP$ is the total pressure on a small area $dF$, we have $$p = \frac{dP}{dF} \qquad \qquad (1)$$ as the pressure per unit area, or intensity of pressure (often, though ambiguously, called the tension in speaking of a gas) on the small surface $dF$. If pressure of the same intensity exists over a finite plane surface of area $F$, the total pressure on that surface is $$P = pFdF = pFdF = Fp,$$ or $$p = \frac{P}{F}. \qquad \qquad (9)$$ (N.B.—For brevity the single word “pressure” will sometimes be used, instead of intensity of pressure, where no ambiguity can arise.) Thus, it is found that, under ordinary conditions at the sea level, the atmosphere exerts a normal pressure (normal to the surface) equal to an intensity of pressure of about $p = 14.7$ lbs per sq. inch (= 2116 lbs per sq. ft.). This intensity of pressure is called one atmosphere. For ex 520 MECHANICS OF ENGINEERING. ample, the total atmospheric pressure on a surface of 100 sq. in. is (inch, lb., sec.) $$P = Fp = 100 \times 14.7 = 1470 \text{ lbs.} \quad (= 0.735 \text{ tons})$$ The quality of $p$ is evidently one dimension of force divided by two dimensions of length. 418. Hydrostatic Pressure: per Unit Area, in the Interior of a Fluid at Rest—In a body of fluid of uniform heaviness, at rest, it is required to find the mutual pressure per unit area between the portions of fluid on opposite sides of any imaginary cutting plane. As customary, we shall consider portions of the fluid as free bodies, by supplying the forces exerted on them by all contiguous portions (of fluid or vessel wall), also those of the earth (their weights), and then apply the conditions of equilibrium. First, cutting plane horizontal.—Fig. 451 shows a body of homogeneous fluid confined in a rigid vessel closed at the top with a small air-tight but frictionless piston (a horizontal disk) of weight $G$ and exposed to atmospheric pressure ($= p_a$ per unit area) on its upper face. Let the area of piston-face be $F$. Then for the equilibrium of the piston the total pressure between its under surface and the fluid at $O$ must be $$P = G + Fp_a,$$ and hence the intensity of this pressure is $$p_i = \frac{G}{F} + p_w \qquad\qquad (1)$$ It is now required to find the intensity, $p$, of fluid pressure between the portions of fluid contiguous to the horizontal cutting plane $BC$ at a vertical distance $h$ vertically below the piston $O$. In Fig. 459 we have as a free body the right parallelo- A diagram showing a horizontal piston in a container filled with fluid. The piston has an area $F$ and is subject to a weight $G$ and atmospheric pressure $p_a$. The diagram includes labels for different parts of the system.

Fig. 451

FLUID PRESSURE. 531 piped $OBC$ of Fig. 451 with vertical sides (two || to paper and four $\uparrow$ to it). The pressures acting on its six faces are normal to them respectively, and the weight of the prism is = vol. $X \times Y$, supposing $Y$ to have the same value at all parts of the column (which is practically true for any height of liquid and for any given height of gas). Hence, if the prism is in equilibrium under the forces shown in the figure, and would still be so were it to become rigid, we may put ($\S 36$) $$Z \text{(vert. comp.)} = 0 \text{ and hence obtain}$$ $$F_P - F_{P_y} - F_{A_Y} = 0 \dots \quad (2)$$ (In the figure the pressures on the ver- tical faces || to paper have no vertical com- ponents, and hence are not drawn.) From (2) we have $$p = p_x + h_Y \quad \quad \quad (3)$$ ($h_Y$, being the weight of a column of homogeneous fluid of unity cross-section and height $h$, would be the total pressure on the base of such a column, if it rest also with no pressure on the upper two faces because that part is acted upon by gravity only.) Secondly, cutting plane oblique—Fig. 452. Consider from an infinitely small right triangular prism $bcd$, whose bases are || to the paper, while the three side faces (rectangles), having areas = $dF_x$, $dF_y$, and $dF_z$, are respectively horizontal, vertical, and oblique; let angle $\alpha = \alpha$. The surface $bc$ is a portion of the plane $BC$ of Fig. 452. Given the intensity of pressure on each face (i.e., $p_x$, $p_y$, and $p_z$), required $p_o$, the intensity of pressure on the oblique face $bd$, of area $dF_o$. (N.B.—The prism is taken very small in order that the intensity of pressure may be considered constant over any one face; and also that the weight of the prism may be neglected, since it involves the volume (three dimen- 522 MECHANICS OF ENGINEERING. sions) of the prism, while the total face pressures involve only two, and is hence a differential of a higher order.] From $Z$ (vert. comp.) = 0 we shall have $$p.dF.cos\alpha - pdF = 0; \text{ but } dF + dF_1 = cos\alpha;$$ $$\therefore p_1 = p_2, \dots \dots \dots (4)$$ which is independent of the angle $\alpha$. Hence, the intensity of fluid pressure at a given point is the same on all imaginary cutting planes containing the point. This is the most important property of a fluid, and is true whether the liquid is at rest or has any kind of motion; for, in case of uniform accelerated motion, e.g., although the sum of the force-components in any direction may be zero, the equation does not in general equal $0$, but = mass × acc.; still, the mass of the body in question is = weight + g, and therefore the term mass × acc. is a differential of a higher order than the other terms of the equation, and hence the same result follows as when there is no motion (or uniform rectilinear motion). 413. The Intensity of Pressure is Equal at all Points of any Horizontal Plane in a body of homogeneous fluid at rest. If we consider a right prism of the fluid in Fig. 451, of small vertical thickness, its axis lying in any horizontal plane $BC$, its bases will be vertical and of equal area $dF$. The pressures on its sides, being normal to them, and hence to the axis, have no components[1] to the axis. The weight of the prism also has no horizontal component. Hence from $Z$ (hor. comp.) = 0, we have, $p_1$ and $p_2$ being the pressure-intensities at the two bases, $$p.dF - p_1.dF = 0; \therefore p = p_1, \dots \dots (1)$$ which proves the statement at the head of this article. It is now plain, from this and the preceding article, that the pressure-intensity $p$ at any point in a homogeneous fluid at rest is equal to that at any higher point plus the weight FLUID PRESSURE. 523 (A-y) of a column of the fluid of section unity and of altitude (A) = vertical distance between the points. i.e., $$p = p_0 + h y, \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \\ p = p_0 + h y; \\ p_4 = p_0 - A y; and p_4 = p_4; \\ hence, finally, by addition we have \\ p_4 = p_0 + h y \\ (in which h = h_i - h_o). \\ If, therefore, upon a small piston at e, of area = F_e, a force P_e be exerted, and an inelastic fluid (liquid) completely fills the vessel, then, for equilibrium, the force to be exerted upon the piston at e, viz., F_e, is thus computed : For equilibrium of fluid $$p_e = p_0 - h y$$ and for equil. of piston at $$p_e = P_e + F_e$$; also, \therefore P_e = F_e P_e + F_e A y. \\ \\ From (8) we learn that if the pistons are at the same level ($$h=0$$) the total pressures on their inner faces are directly proportional to their areas. If the fluid is gaseous (9) and (8) are practically correct if A is not > 100 feet (for gas being compressible, the lower strata are generally more dense than the upper), but in (3) the pistons must be fixed, and $$P_e$$ and $$F_e$$ refer solely to the interior pressures. 624 MECHANICS OF ENGINEERING. Again, if $h$ is small or $p$, very great, the term $k\gamma$ may be omitted altogether in eqs. (2) and (3) (especially with gases, since for them $\gamma$ (heaviness) is usually small), and we then have, from (2), $$p = p_i \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \dots \\ being the algebraic form of the statement: A body of fluid at rest transmits pressure with equal intensity in every direction and to all of its parts. [Principle of "Equal Transmission of Pressure."]$ 414. Moving Pistons.—If the fluid in Fig. 454 is inelastic and the vessel walls rigid, the motion of one piston ($o$) through a distance $s$, causes the other to move through a distance $s$, determined by the relation $F_o = F_p$ (since the volumes described by them must be equal, as liquids are incompressible); but on account of the inertia of the liquid, and friction on the vessel walls, equations (2) and (3) no longer hold exactly, still are approximately true, though they become less so as the vessel short, as with the cylinder of a water-pressure engine. But if the fluid is compressible and elastic (gases and vapors; steam, or air) and hence of small density, the effect of inertia and friction is not appreciable in short wide vessels like the cylinders of steam- and air-engines, and those of air-compressors; and eqs. (2) and (3) still hold practically, even with high piston-speeds. For example, in the space $A B$, Fig. 455, between the pistons and cylinder-head $A$, when the moving piston $B$ (moving toward the right) the intensity of pressure against the steam against the moving piston $B$ is practically equal to that against the cylinder-head $A$ at the same instant. 415. An Important Distinction between gases and liquids (i.e., between elastic and inelastic fluids) consists in this: A liquid can exert pressure against the walls of the containing vessel only by its weight, or (when confined on all sides) by transmitted pressure coming from without (due to piston pressure, atmospheric pressure, etc.), whereas— FLUID PRESSURE. 095 A gas, confined, as it must be, on all sides to prevent diffusion, exerts pressure on the vessel not only by its weight, but by its elasticity or tendency to expand. If pressure from without is also applied, the gas is compressed and exerts a still greater pressure on the vessel walls. 416. Component of Pressure, in a Given Direction.—Let $ABCD$, whose area = $dF$, be a small element of a surface, plane or curved, and $p$ the intensity of fluid pressure upon this element, then the total pressure upon it is $pdF$, and is of course normal to it. Let $A'B'CD'$ be the projection of the element $DF$ upon a plane $CDM$, and let $\alpha$ be the angle between the element, and let it be required to find the value of the component of $pdF$ in a direction normal to this last plane (the other component being understood to be in the same plane). We shall have Component of $pdF \cdot \gamma$ to $CDM = pdF\cos \alpha = p(dF.\cos \alpha)$. (1) But $dF.\cos \alpha = area A'B'CD'$, the projection of $dF$ upon the plane $CDM$; :. Component $\gamma$ to plane $CDM = p \times (\text{project. of } dF \text{ on } CDM)$; i.e., the component of fluid pressure (on an element of a surface) in a given direction (the other component being $\gamma$ to the first) is found by multiplying the intensity of the pressure by the area of the projection of the element upon a plane $\gamma$ to the given direction. It is seen, as an example of this that if the fluid pressures on the elements of this inner surface of one hemisphere of a hollow sphere containing a gas are resolved into components and $\gamma$ to the plane of the circular base of the hemisphere, the sum of the former components simply = $sr^2p$, where $r$ is the radius of the sphere, and $p$ the intensity of the fluid pressure; for, from the foregoing, the sum of these components is just the same as the total pressure would be, having an intensity $p$, Fig. 43. Fig. 43. 526 MECHANICS OF ENGINEERING. on a great circle of the sphere, the area, $\pi r^2$, of this circle being the sum of the areas of the projections, upon this circle as a base, of all the elements of the hemispherical surface. (Weight of fluid neglected.) A similar statement may be made as to the pressures on the inner curved surface of a right cylinder. 417. Non-planar Pistons.—From the foregoing it follows that the sum of the components $|p|$ of the piston-rod, of the fluid pressure acting on the plane $\gamma$ (Fig. 457), is just the same as at $R$, if the cylinders are of equal length, diameter, or area, is at the same tension. For the sum of the projections of all the elements of the curved surface of $A$ upon a plane $\gamma$ to the piston-rod is always $\pi r^2 = \text{area of section of cylinder-bore}$. A diagram showing a cross-section of a piston rod and cylinder with a curved surface. The diagram shows two different views: one from above and one from the side. Fig. 457. If the surface of $A$ is symmetrical about the axis of the cylinder, then other components (i.e., those $\gamma$ to the piston-rod) will neutralize each other. If the line of intersection of that surface with the surface of the cylinder is not symmetrical about the axis of the cylinder, the piston may be pressed laterally against the cylinder by means of the thrust along the rod or over-hanging (§ 128) is the result. In this case, friction induced by the lateral pressure, in all instances, as if the surface were plane and $\gamma$ to piston-rod. 418. Bramah or Hydraulie Press.—This is a familiar instance of the principle of transmission of fluid pressure. Fig. 458. Let the small piston at $O$ have a diameter $d = 1$ inch = $\frac{1}{12}$ ft., while the plunger $E$, or large piston, has a diameter $d' = AB = CD = 15$ in. = $\frac{1}{4}$ ft. The lever $MN$ weighs $G_1 = 3$ lbs., and a weight $G = 40$ lbs. is hung at $M$. The lever-arma of these forces about the fulcrum $N$ are given in the figure. The apparatus being full of water (oil is often used), the fluid pressure $P$, against the small piston is found by putting $$\text{Area} \times P = \text{Weight} \times \text{Lever-Arm}$$ $$\frac{\pi d^2}{4} \times P = G_1 \times MN$$ $$\frac{\pi d'^2}{4} \times P = G \times MN$$ Solving for $P$, we get: $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = \frac{G_1 \times MN}{\frac{\pi d^2}{4}} = \frac{G \times MN}{\frac{\pi d'^2}{4}}$$ $$P = ... FLUID PRESSURE. 527 $F_{\text{(moms. about } N)} = 0 \text{ for the equilibrium of the lever;}$ whence [ft., lb., sec.] $$P_s \times 1 - 40 \times 3 - 3 \times 9 = 0. \therefore P_s = 196 \text{ lbs.}$$ Diagram showing a lever with weights and a piston. **Fig. 68** But, denoting atmospheric pressure by $p_a$, and that of the water against the piston by $p_s$ (per unit area), we may also write $$P_s = F_s p_s - F_p p_a = \frac{1}{4} w d (p_s - p_a).$$ Solving for $p_s$, we have, putting $p_a = 14.7 \times 144$ lbs. per sq. ft. $$p_s = \left[196 + \frac{\pi}{4} (\frac{d}{4})^2\right] + 14.7 \times 144 = 25236 \text{ lbs. per sq. ft.}$$ Hence at $e$ the press. per unit area, from § 409, and (2), § 413, is $$p_e = p_s + h y = 25236 + 3 \times 62.5 = 25423 \text{ lbs. per sq. ft.}$$ = 175.6 lbs. per sq. inch or 11.9 atmospheres, and the total upward pressure at $e$ on base of plunger is $$P = F_e p_e = \pi \frac{d^2}{4} p_s = \frac{1}{4} \pi (\frac{d}{4})^2 \times 25423 = 21194 \text{ lbs.,}$$ or almost 16 tons (of 2000 lbs. each). The compressive force upon the block or bale, $C$, = $P$ less the weight of the plunger and total atmos. pressure on a circle of 15 in. diameter. 528 MECHANICS OF ENGINEERING. 418. The Dividing Surface of Two Fluids (which do not mix) in Contact, and at Rest, is a Horizontal Plane. —For, Fig. 459, supposing any two points $a$ and $O$ of this surface to be at different levels (the pressure at $O$ being $p_{o}$, that at $p_{a}$, and the heaviest of the two fluids $\gamma$, and $\gamma'$ respectively), we shall have the following configuration of the two elementary prisms $ab$ and $bo$ (vertical and horizontal), the relation $$p_a = p_o + h\gamma'$$ while from the prisms $ac$ and $co$, the relation $$p_c = p_o + h\gamma.$$ These equations are conflicting, hence the above supposition is absurd. Therefore the proposition is true. For stable equilibrium, evidently, the heavier fluid must occupy the lowest position in the vessel, and if there are several fluids (which do not mix), they will arrange themselves vertically, in order of their densities, the heaviest at the bottom, Fig. 460. On account of the property called diffusion the particles of two gases placed in contact soon intermingle and form a uniform mixture. This fact gives strong support to the "Kinetic Theory of Gases" (§ 408). 420 Free Surface of a Liquid at Rest.—The surface (of a liquid) not in contact with the walls of the containing vessel is called a free surface, and is necessarily horizontal (from § 419) when the liquid is at rest. Fig. 461. (A gas, from its tendency to indefinite expansion, is incapable of having a free surface.) This is true even if the space between the liquid and the wall for if the surface were inclined or curved, points in the body of the liquid and in the same horizontal plane would have different heights (or "heads") of liquid. Fig. 460 Fig. 461 TWO LIQUIDS IN BENT TUBE. 599 between them and the surface, producing different intensities of pressure in the plane, which is contrary to § 413. When large bodies of liquid like the ocean are considered, gravity can no longer be regarded as acting in parallel lines; consequently the free surface of such a liquid is curved, being $\gamma$ to the direction of (apparent) gravity at all points. For ordinary engineering purposes (except in Goodby's free surface of water at rest is a horizontal plane. 413. Two Liquids (which do not mix) at Rest in a Bent Tube open at Both Ends to the Air, Fig. 460; water and mercury, for instance. Let their heavinesses be $y_{1}$ and $y_{2}$ respectively. The pressure at e may be written (§ 413) either $$p_e = p_a + h_1 y_1 \dots (1)$$ or $$p_e = p_a + h_2 y_2 \dots (2)$$ according as we refer it to the water column or the mercury column and their respective free surfaces where the pressure $p_a = p_b = p_c = \text{atmos. press.}$. e is the surface of contact of the two liquids. Hence we have $$p_a + h_1 y_1 = p_a + h_2 y_2 \dots i.e., h_1 : h_2 :: y_1 : y_2 \dots (3)$$ i.e., the heights of the free surfaces of the two liquids above the surface of contact are inversely proportional to their respective heavinesses. Example 1. If the pressure at e = 2 atmospheres (§ 390) we shall have from (3) (inch-lib.-sec. system of units) $$h_1 y_1 = p_e - p_a = 3 \times 14.7 - 14.7 = 14.7 \text{ lbs. per sq. inch.}$$ $$\therefore h_1 \text{ must } = 14.7 + [848.7 + 1728] = 30 \text{ inches}$$ (dince, for mercury, $y_2 = 848.7$ lbs. per cub. ft.). Hence, from (3), $$h_1 y_1 = \frac{30 \times [848.7 + 1728]}{62.5 + 1728} = 408 \text{ inches} = 34 \text{ feet}.$$ 530 MECHANICS OF ENGINEERING. i.e., for equilibrium, and that $p_1$ may = 9 atmospheres, $h_1$ and $h_2$ (of mercury and water) must be 30 in. and 34 feet respectively. 429 City Water-pipes.—If $A$ = vertical distance of a point $B$ of a water-pipe below the free surface of reservoir, and the water be at rest, the pressure on the inner surface of the pipe at $B$ (per unit of area) is $$p = p_1 + h_2; \text{ and here } p_1 = p_a = \text{atmos. press}.$$ **Example.**—If $h = 200$ ft. (using the inch, lb., and second) $$p = 14.7 + [200 \times 12] (62.5 - 1728) = 101.5 \text{ lbs. per sq. in.}$$ The term $h_2$, alone = 98.8 lbs. per sq. inch, is spoken of as the **hydrostatic pressure** due to 200 feet height, or "Head," of water. (See Troutwine's Pocket Book for a table of hydrostatic pressures for various depths.) If, however, the water is flowing through the pipe, the pressure against the interior wall becomes less (a problem of Hydrodynamics to be treated subsequently), while if that motion is suddenly checked, the pressure becomes momentarily much greater than the hydrostatic. This shock is called "water-rum" and "water-hammer," and may be as great as 200 to 300 lbs. per sq. inch. 423 Barometers and Manometers for Fluid Pressure.—If a tube, closed at one end, is filled with water, and the other extremity is temporarily stopped and afterwards opened to the atmosphere, the water will rise then (vertical) higher than above its surface of the water, it is required to find the intensity, $p_1$, of fluid pressure at the top of the tube, supposing it to remain filled with water. Fig. 463. At $E'$ inside the tube the pressure is 14.7 lbs. per sq. inch, the same as that outside at the same level ($\S$ 413); hence, from $p_2 = p_1$ $$p_1 = p_E - h_2 \dots \dots \dots \dots \dots (1')$$ **BABOMETERS.** EXAMPLE.—Let $h = 10$ feet (with inch-lb.-sec. system); then $$p_s = 14.7 - 120 \times (02.5 + 1728) = 10.4 \text{ lbs. per sq. inch},$$ or about $p_s$ of an atmosphere. If we now inquire the value of $A h$ to make $p_s = 0$, we put $p_s - A h = 0$ and obtain $A = 408$ inches, = 34 ft., which is called the height of the water-barometer. Hence, Fig. 463a, ordinary atmospheric pressure will not sustain a column of water higher than 34 feet. If mercury is used instead of water the height supported by one atmosphere $$b = 14.7 + [848.7 + 1728] = 30 \text{ inches},$$ = 76 centima (about), and the tube is of more manageable proportions than with water, aside from the advantage that no vapor of mercury forms above the liquid at ordinary temperatures. [In fact, the water-barometer height $b = 34$ feet has only a theoretical existence since at ordinary temperatures ($40^\circ$ to $80^\circ$ Fahr.) vapor of water would form above the column and depress it by from 0.30 to 1.09 ft.] Such an apparatus is called a Barometer, and is used not only for measuring the varying variations in the barometric pressure but also to indicate the weather and height above sea-level, but also that of any body of gas. Thus, Fig. 464, the gas in $D$ is put in communication with the space above the mercury in the cistern at $C$; and we have $p = h y$, where $y$ is heat of mercury, and $p$ is the pressure on the liquid in the cistern. For delicate measurements an attached thermometer is also used, as the heat-motion of the gas may affect this temperature. If the vertical distance $CD$ is small, the tension in $C$ is considered the same as in $D$. For gas-tensions greater than one atmosphere, the tube may be left open at the top, forming an open mas- 532 MECHANICS OF ENGINEERING. nometer, Fig. 465. In this case, the tension of the gas above the mercury in the cistern is $$p = (h + b)y,$$ .......................... (1) in which $b$ is the height of mercury (about 30 in.) to which the tension of the atmosphere above the mercury column is equivalent. EXAMPLE.—If $A = 51$ inches, Fig. 465, we have $(ft., lb., sec.)$ $$p = [4.25 ft. + 2.5 ft.] \times 848.7 = 5728 \text{ lbs. per sq. foot}$$ $$= 39.7 \text{ lbs. per sq. inch} = 9.7 \text{ atmospheres}.$$ Another form of the open manometer consists of a U tube, Fig. 464, the atmosphere having access to one branch, the gas to be examined, to the other, while the mercury lies in the curve. As before, we have $$p = (h + y)y = h y + p_a,$$ .......................... (2) where $p_a =$ atmos. tension, and $y$ as above. The tension of a gas is sometimes spoken of as measured by so many inches of mercury. For example, a tension of 29.05 lbs. per sq. inch (14 atmos.) is measured by 45 inches of mercury in a vacuum manometer (i.e., a common barometer), Fig. 464. With the open manometer this tension (14 atmos.) would be indicated by 15 inches of actual mercury, Figs. 465 and 466. An ordinary steam-gauge indicates the excess of tension over one atmosphere ; thus "40 lbs. of steam" implies a tension of $40 + 14.7 = 54.7$ lbs. per sq. in. The Bourdon-tube-manometer is used in place of a curved elastic metal tube of flattened or elliptical section (with the long axis $\gamma$ to the plane of the tube), and has one end fixed. The movement of the other end, which is free and A diagram showing a Bourdon-tube manometer with a curved elastic metal tube. TENSION OF GASES. 533 closed, by proper mechanical connection gives motion to the pointer of a dial. This movement is caused by any change of tension in the steam or gas admitted, through the fixed end, to the interior of the tube. As the tension increases the elliptical section becomes less flat, i.e., more nearly circular, causing the two ends of the tube to separate more widely, i.e., the free end moves away from the fixed end; and vice versa. Such gauges are called "mercury barometers," but they are graduated for comparison with mercury manometers; and should be tested from time to time in the same way.* 424. Tension of Illuminating Gas.--This is often spoken of as measured by inches of water (from 1 to 3 inches usually). Strictly it should be stated that this water-height measures the excess of its tension over that of the atmosphere. Thus, in Fig. 466, water being used instead of mercury, $A = \text{say} \ 2$ inches, while $\delta = 408$ inches. This difference of tension may be largely affected by a change in the barometer due to the weather, or by a difference in altitude, as the follow- ing example will illustrate: Example.--Let us take the gas at rest, and the tension at the geometer $A$, Fig. 467, to be "two inches of water," required the water-column $h'$ (in open tube) that the gas will support in the pipe at $B$, 120 feet (vertically) above the geometer. Let the temperature be freezing (nearly), and the outside air at a tension of 14.7 lbs. per sq. inch; the heaviness of the gas at this temperature being 0.038 lbs. per cubic foot. For the small difference of 120 ft. we may treat both the atmosphere and the gas as liquids, that is, constant density throughout the vertical column, and therefore apply the principles of § 413; with the following result: The tension of the outside air at $B$, supposed to be at the same temperature as at $A$, will sustain a water-column less than the 408 inches at $A$ by an amount corresponding to the *Of late years gauges have come into use constructed of boxes with corrugated sides of thin metal like the aneroid barometer. Motion of the edge, under varying internal fluid pressure, causes movement of a pointer on a dial. A diagram showing a closed system with a movable piston inside a U-tube containing a liquid (likely water) connected to a larger U-tube filled with another liquid (likely mercury). The smaller U-tube has a scale marked in inches along one arm. Fig. 466 534 MECHANICS OF ENGINEERING. 190 feet of air between, of the heaviness 0.807 lbs. per cub. ft. 190 feet of air weighing 0.807 lbs. per cub. ft. will balance 0.154 ft. of water weighing 62.5 lbs. per cubic ft., i.e., 1.85 inches of water. Now the tension of the gas at $B$ is also less than its tension at $A$, but the difference is not so great as with the outside air, for the 190 ft. of gas is lighter than the 190 ft. of air. Since 190 ft. of gas weighing 0.086 lbs. per cubic ft. will balance 0.0691 ft., or 0.83 inches of water, therefore the difference between the tensions of the two fluids at $B$ is greater than at $A$ by $(1.85 - 0.83 = )1.02$ inches; or, at $B$ the total difference is greater than at $A$. Hence if a small aperture is made in the pipe at $B$ the gas will flow out with greater velocity than at $A$. At Ithaca, N.Y., where the University buildings are 400 ft. above the gas-works, this phenomenon is very marked. When the difference of level is great the decrease of tension as we proceed upward in the atmosphere, even with constant temperature, does not follow the simple law of § 413; see § 417. For velocity of flow of gases through orifices, see § 548, etc. 425 Safety-valves—Fig. 468. Required the proper weight $G$ to be hung at the extremity of the horizontal lever $AB$, with fulcrum at $B$, that the flat disk-valve $E$ shall not be forced upward by any pressure $p$, until the latter reaches a given value $= p$. Let the weight of the arm be $G_1$, its centre of gravity being at $C$, a distance = $s$ from $B$; the other horizontal distances are marked in the figure. Suppose the valve on the point of rising; then the forces acting on the lever are the fulcrum-reaction at $B$, the weights $G$ and $G_1$, and the two fluid-pressure on the disk, viz., $F_{Pa}$ (atmospheric) downward, and $Fp$ (steam) upward. Hence, $$Ob + Gc + F_{Pa} - Fpa = 0.$$ (1) A diagram showing a lever system with a flat disk-valve E attached to it. BURSTING OF PIPES. 535 Solving, we have $$G = \frac{a}{b} F(p - p_a) - G_0^{\frac{c}{d}} \dots \dots (2)$$ **EXAMPLE.**—With $a = 2$ inches, $b = 2$ feet, $c = 1$ foot, $G_0 = 4$ lbs., $p = 6$ atmos., and diam. of disk = 1 inch; with the foot and pound, $$G = \frac{2}{24} \cdot \frac{\pi(14)}{12} [6 \times 14.7 \times 144 - 1 \times 14.7 \times 144] - 4 \times 8$$ $$\therefore G = 2.81 \text{ lbs.}$$ [Notice the cancelling of the $144$, for $F(p - p_a)$ is pounds, being one dimension of force, if the pound is selected as the unit of force, whether the inch or foot is used in both factors.] Hence when the steam pressure has risen to 6 atmos. (= 88.3 lbs. per square inch) (corresponding to 73.5 lbs. per sq. in., by steam-gauge) the valve will open if $G = 2.81$ lbs., or be on the point of opening. **426. Proper Thickness of Thin Hollow Cylinders (i.e., Pipes and Tubes) to Resist Bursting by Fluid Pressure.** **Case I. Stresses in the cross-section due to End Pressure; Fig. 469.—Let $AB$ be the circular cap closing the end of a cylindrical tube containing fluid at a temperature $T$. Let $r$ denote radius of the tube or pipe. Then considering the cap free, neglecting its weight, we have three sets of $\|j$ forces in equilibrium in the figure, viz.: the internal fluid pressure = $\pi r p$; the external fluid pressure = $\pi r p_a$; while the total stress (tensile) on the small ring, whose area now exposed is $2\pi r t$ (see fig.), is $= \frac{\pi r p}{t}$, where $t$ is the thickness of the pipe, and $p$, the tensile stress per unit area induced by the end-pressure (fluid). A diagram showing a cylindrical tube with a circular cap at one end, labeled "End Pressure." The diagram includes various lines and annotations indicating stresses and pressures. 536 MECHANICS OF ENGINEERING. For equilibrium, therefore, we may put $\Sigma(\text{hor. comp.}) = 0$; i.e., $$\pi r^2 p - \pi r^2 p_a - 2\pi r r y_1 = 0;$$ $$\therefore p_1 = \frac{r(p - p_a)}{2r} \qquad \qquad (1)$$ (Strictly, the two circular areas sustaining the fluid pressures are different in area, but to consider them equal occasions but a small error.) Eq. (1) gives the tension in the central section of a thin hollow sphere, under bursting pressure. CASE II. Stress in the longitudinal section of pipe, due to radial fluid pressures.* Consider free the half (semi-circular) of any length $l$ of the pipe, between two cross-sections. Take an axis $X$ (as in Fig. 470) $\parallel$ to the longitudinal section which has been made. Let $p_i$ denote the tensile stress (per unit area) produced in the narrow angles at $A$ and $B$ (those in the half-ring edges, having no $X$ components, are not drawn in the figure). On the internal curved surface the fluid pressure is considered of equal intensity $= p$ at all points (practically true even with liquids, if $2r$ is small compared with the head of water producing $p$). The fluid pressure on the external curved surface is $= pdF$. Its $X$ component (see § 416) is obtained by multiplying $p$ by the projection of $dF$ on the vertical plane $ABC$, and since $p$ is the same for all the $dF's$ of the curved surface, the sum of all the $X$ components of the internal fluid pressure must = $p$ multiplied by the area of rectangle $ABCD$, = $2rp$; and similarly the $X$ components of the A diagram showing a pipe with a semi-circular section. The pipe is labeled "Pipe" and has two cross-sections labeled A and B. The axis X is shown running through both sections. The diagram also shows a small circle labeled "tension" and a line labeled "internal pressure". * Analytically this problem is identical with that of the smooth cord on a smooth cylinder, § 108, and is seen to give the same result. BURSTING OF PIPES. 597 external atmos. pressure = 2r$p_a$ (nearly). The tensile stresses ($\frac{1}{l}$ to $x$) are equal to $2r$p, ; hence for equilibrium, $2X=0$ gives $$2rp_x - 2rp + 2rp_a = 0;$$ $$\therefore p_x = \frac{r(p-p_a)}{t} \dots \dots \dots \dots \dots (3)$$ This tensile stress, called hoop tension, $p_x$, opposing rupture by longitudinal tearing, is seen to be double the tensile stress $p$, induced, under the same circumstances, on the annular cross-section in Case I. Hence eq. (3), and not eq. (1), should be used to determine a safe value for the thickness of metal, $t$, or any other one unknown quantity involved in the equation. For safety against rupture, we must put $p_x = T'$, a safe tensile stress per unit area for the material of the pipe or tube (see §§ 195 and 205). $$\therefore t = \frac{r(p-p_a)}{T'} \dots \dots \dots \dots \dots (3)$$ (For a thin hollow sphere, $t$ may be computed from eq. (1); that is, need be only half as great as with the cylinder, other things being equal.) EXAMPLE.-A pipe of twenty inches internal diameter is to contain water at rest under a head of 340 feet; required the proper thickness, if of cast-iron. 340 feet of water measures 10 atmospheres, so that the internal fluid pressure is 11 atmospheres; but the external pressure $p_a$ being one atmos., we must write (inch, lb., sec.) $$(p-p_a) = 10 \times 14.7 = 147.0 \mathrm{lbs.~per~sq.~in.,~and~} r = 10 \mathrm{in.,}$$ while (§ 908) we may put $T' = \frac{1}{4}$ of 9000 = 4500 lbs. per sq. in.; whence $$t = \frac{10 \times 147}{4500} = 0.336 \mathrm{inches.}$$ 588 MECHANICS OF ENGINEERING. But to insure safety in handling pipes and imperviousness to the water, a somewhat greater thickness is adopted in practice than given by the above theory. Thus, Weibach recommends (as proved experimentally also) for
Pipes of sheet iron, $t = [0.00172 rA + 0.13]$ inches ;
cast iron $t = [0.00476 rA + 0.34]$"
" copper $t = [0.00296 rA + 0.16]$"
" lead $t = [0.01014 rA + 0.21]$"
" zinc $t = [0.00484 rA + 0.16]$"
in which $t =$ thickness in inches, $r =$ radius in inches, and $A =$ excess of internal over external fluid pressure (i.e., $p - p_s$) expressed in atmospheres. For instance, for the example just given, we should have (cast-iron) $t = .00476 \times 10 \times 10 + 0.34 = 0.816$ inches. If the pipe is subject to "water-ram" ($\S$ 429) the strength should be much greater. To provide against "water-ram," Mr. J. T. Fanning, on p. 453 of his "Hydraulic and Water-supply Engineering," advises adding 230 feet to the static head in computing the thickness of cast-iron pipes. For thick hollow cylinders see Rankine's Applied Mechanics, p. 290, and Cotterill's Applied Mechanics, p. 468. 497. Collapsing of Tubes under Fluid Pressure. (Cylindrical boiler-flues, for example.)--If the external exceeds the internal fluid pressure, and the thickness of metal is small compared with the diameter, the slightest deformation of the tube or pipe gives the external pressure greater capability to produce a further change of form, and hence possibly a final collapse; just as a thin wire will bend a slight angle without giving great advantage to the terminal force. Hence the theory of § 426 is inapplicable. According to Sir Wm. Fairbairn's experiments (1858) a thin wrought-iron cylindrical (circular) tube will not collapse until the excess of external over internal pressure is COLLAPSE OF TUBES. 539 $$p(\mathrm{in}\,\mathrm{lb.\ per\ sq.\ in.})=9672000\frac{d}{l^2} \cdot \cdot \cdot (1) \cdot \cdot \cdot (\mathrm{not\ homogeneous})$$ (e., l. and d must all be expressed in the same linear unit.) Here $t =$ thickness of the wall of the tube, $d$ its diameter, and $l$ its length, the ends being understood to be so supported as to preclude a local collapse. **EXAMPLE.**—With $l = 10$ ft. = 190 inches, $d = 4$ in., and $t = \frac{1}{8}$ inch, we have $$p = 9672000\left[\frac{1}{100} \div (190 \times 4)\right] = 201.5\,\mathrm{lbs.\ per\ sq.\ inch}.$$ For safety, $\frac{4}{3}$ of this, viz. 40 lbs. per sq. inch, should not be exceeded; e.g., with 14.7 lbs. internal and 54.7 lbs. external. [Note.—For simplicity the original of the thickness used in eq. (1) above has been omitted. In the original formula it is $5$ and these dimensions must be expressed in terms of $l$. A discussion of the elements of Mr. Fairbairn will be found in a paper read by Prof. Unwin before the Institute of Civ. Engineers(Proceedings, vol. xvi). See also Prof. Cowin's "Machine Design," p. 288.] In practice, however, owing to the various conditions of service, boiler-flues are subjected to such serious straining actions due to unequal expansion of the connecting parts as to render the above formula quite unreliable, thus requiring a large allowance in its application.] 540 MECHANICS OF ENGINEERING. CHAPTER II. HYDROSTATICS (Continued)---PRESSURE OF LIQUIDS IN TANKS AND RESERVOIRS. 428. Body of Liquid in Motion, but in Relative Equilibrium.--By relative equilibrium it is meant that the particles are not changing their relative positions, i.e., are not moving among each other. On account of this relative equilibrium the following problems are placed in the present chapter, instead of under the head of Hydrodynamics, where they strictly belong. As regards the liquid, owing to its elastic property of rigid bodies, we may apply the equations of motion of rigid bodies to bodies of liquid in relative equilibrium. CASE I. All the particles moving in parallel right lines with equal velocities; at any given instant (i.e., a motion of translation.)--If the common velocity is constant we have a uniform translation, and all the forces acting on any one particle are balanced, as if it were not moving at all (according to Newton's Laws, § 64). Hence the relative motion of the particles, free motions, will be such that the surface of the liquid will be at rest. Thus, Fig. 471. if the liquid in the moving tank is at rest relatively to the tank at a given instant, with its free surface horizontal, and the motion of the tank be one of translation with a uniform velocity, the liquid will remain in this condition of relative rest, as the motion proceeds. But if the velocity of the tank is accelerated with a constant acceleration $p = p$ (this symbol must not be confused with $p$ for pressure), the free surface will begin to oscillate, and finally come to relative equilibrium at some angle $\alpha$ with the horizontal, which is thus found, when the motion is horizontal. See Fig. 472, in which the position and value of $\alpha$ are the same whether the motion is uniformly accelerated from left to right. Fig. 471. Fig. 472. RELATIVE EQUILIBRIUM OF LIQUIDS. 541 or uniformly retarded from right to left. Let $O$ be the lowest point of the free surface, and $Ob$ a small prism of the liquid with its axis horizontal, and of length = $x$; $nb$ is a vertical prism of length = $z$, and extending from the extremity of $Ob$ to the free surface. The pressure at both $O$ and $n$ is $p_a=$ atmos. pres. Let the area of cross-section of both prisms be = $dF$. Now these forces acting in direction $X$ (horizontal), the difference between the forces on its two ends, i.e., its $\vec{X} \vec{F}$ must = its mass $\times$ accel. (§ 109). $\therefore p_d dF - p_d dF = (xdF + yg) \vec{p} \cdot \cdot \cdot (1)$ ($y =$ heaviness of liquid; $p_b =$ press. at $b$; and since the vertical prism $nb$ has no vertical acceleration, the $\vec{X}$ (vert. compo.) for it must = 0. $\therefore p_d dF - p_d dF - sdF$. $y = 0 \cdot \cdot \cdot (2)$ From (1) and (2), $\frac{xy}{g} \vec{p} = xy; \therefore \frac{x}{a} = \frac{\vec{p}}{g} \cdot \cdot \cdot (3)$ Hence $On$ is a right line, and therefore $\tan \alpha$, or $\frac{x}{a} = \frac{\vec{p}}{g} \cdot \cdot \cdot (4)$ [Another, and perhaps more direct, method of deriving this result is to consider first a small particle of the liquid lying in the surface. The force acting on this particle is two: the first its weight = $dF$; and the second the resultant action of its immediate neighbor-particles. Now this latter force (pointing obliquely upward) must be normal to the free surface of the liquid, and therefore must make the unknown angle $\alpha$ with the vertical. Since the particle has at this instant a rectilinear accelerated motion in a horizontal direction, the resultant of the two forces mentioned must be horizontal and have a value = mass $\times$ acceleration. That is, the diagonal formed on the two] A diagram showing two prisms of liquid, one horizontal ($Ob$) and one vertical ($nb$), with their respective pressures indicated. Para. 67 542 MECHANICS OF ENGINEERING. forces must be horizontal and have the value mentioned, = $(dG + q)g$; while from the nature of the figure (let the student make the diagram for himself) it must also $= dG \tan a$. $\therefore dG \tan a = \frac{dG}{g} p$; or, $\tan a = \frac{\overline{z}}{g}$ . . . . . Q. E. D.] If the translation were vertical, and the acceleration upward [i.e., if the vessel had a uniformly accelerated upward motion or a uniformly retarded downward motion], the free surface would be horizontal, but the pressure at a depth $= h$ below the surface would be $p - h g$, and $h g$ would be obtained by calculus. Considering free a small cylindrical prism of height $= A$ with upper base in the free surface, and putting $2(vert.\ comp.) = mass \times acceleration$, we have $$dF.p - dF.p_a - hdF.y = \frac{hdF.y}{g}. \bar{p};$$ $\therefore p = p_a + h y \left[ \frac{q+p}{g} \right]$. . . . . (5) If the acceleration is downward (not the velocity necessarily) we make $p$ negative in (5). If the vessel falls freely, $p = - g$ and $\therefore p = p_a$, in all parts of the liquid. Query : Suppose $p$ downwards and > $g$. **Case II. Uniform Rotation about a Vertical Axis.** If the narrow vessel in Fig. 473, open at top and containing a liquid, be kept rotating at a uniform speed about a vertical axis $Z$, the liquid after some oscillations will be brought (by friction) to relative equilibrium (rotating about $Z$, as if rigid). Required the form of the free surface (evidently a surface of revolution) at each point of which we know $p$. Let $O$ be the intersection of the axis $Z$ with the surface, and $n$ any point in the surface; $b$ being A diagram showing a narrow vessel with a free surface rotating about a vertical axis Z. Fig. 473 UNIFORM ROTATION OF LIQUID IN VESSEL. 543 a point vertically under $n$ and in same horizontal plane as $O$. Every point of the small right prism $nh$ (of altitude = $z$ and sectional area $dF$) is describing a horizontal circle about $Z$, and has therefore no vertical acceleration. Hence for this prism, free, we have $\Sigma Z = 0$; i.e., $$dF.p_n - dF.p_a - sdF.y = 0 \dots (1)$$ Now the horizontal right prism $Ob$ (call the direction $O \ldots b$, $X$) is rotating uniformly about a vertical axis through one extremity, as if it were a rigid body. Hence the forces acting on it must be equivalent to a single horizontal force, $- \omega^2 M_p$ ($\S 122a$), coinciding in direction with $X$. ($M =$ mass of prism = its weight + $g$ and $p =$ distance of its centre of gravity from $O$; here $p = \frac{1}{2} \frac{1}{2}$ length of prism). Hence the $\Sigma X$ of the forces acting on the prism $Ob$ must = $- \omega^2 M_p \cdot g \cdot y$. But the forces acting on the two ends of this prism are their own $X$ components, while the lateral pressures and the weights of its particles have no $X$ components. $$\therefore dF.p_a - dF.p_b = - \frac{\omega^2 dF.y}{2g} \dots (2)$$ From (1) and (2) we have $$z = \frac{(wx)^2}{2g} = \frac{v^2}{2g}; \dots (3)$$ where $v = ax =$ linear velocity of the point $n$ in its circular path. [As in Case I, we may obtain the same result by considering a single surface particle free, and would derive for the resultant force acting upon it the value $dF\tan x$ in a horizontal direction and intersecting the axis of rotation. But here $x$ is different for particles at different distances from the axis, $\tan x$ being the $\frac{dx}{dy}$ of the curve $On$. As the particle is moving uniformly in a circle the resultant force must point toward the] 544 MECHANICS OF ENGINEERING. centre of the circle, i.e., horizontally, and have a value $\frac{dG}{z} \cdot z'$, where $z$ is the radius of the circle [§ 74, eq. (5)]: $\therefore dG \tan \alpha = \frac{dG}{g} \cdot z' \text{ or } \tan \alpha = \frac{dz}{dx} = \frac{dw}{dx}$ $\therefore \int_0^{\infty} dz = \frac{dG}{g} \int_0^{\infty} dx; \text{ or, } z = \frac{dG}{g} \cdot Q.E.D.$ Hence any vertical section of the free surface through the axis of rotation $Z$ is a parabola, with its axis vertical and vertex at $O$; i.e., the free surface is a paraboloid of revolution, with $Z$ as its axis. Since $w$ is the linear velocity $v$ of the point $b$ in its circular path, $s = "height due to velocity" v [§ 52]. EXAMPLE.--If the vessel in Fig. 473 makes 100 revolutions per minute, required the ordinate $z$ at a horizontal distance of $x = 4$ inches from the axis (ft.-lb.-sec. system). The angular velocity $\omega = (2\pi 100 + 60) / 3600$ radians per sec. [N.B.--A radian = the angle subtended by an arc equal to the radius]. At half-revol, or angle of $180^\circ$. With $x = 4$ ft. and $g = 32.2$, $s = \frac{d^2z}{dy} = \left(\frac{10\pi}{3}\right)^2 \left(\frac{1}{3}\right)^2 \frac{1}{64.4} = 0.188 \text{ ft.} = 9\frac{1}{4} \text{ inches},$ and the pressure at $b$ (Fig. 471) is (now use inch, lb., sec.) $p_b = p_a + gy = 14.7 + 9\frac{1}{4} \times 69.5 = 14.781 \text{ lbs. per sq. in.}$ Prof. Mendelejeff of Russia has recently utilized the fact announced as the result of this problem, for forming perfectly true paraboloidal surfaces of plaster of Paris, to receive by galvanic process a deposit of metal, and thus produce specula of exact figure for reflecting telescopes. The vessel containing the liquid plaster is kept rotating about a vertical axis at the proper uniform speed, and the plaster assumes the desired shape and form on the inner surface. A similar vessel, filled, may also be placed in the vessel, instead of liquid plaster. **RELATIVE EQUILIBRIUM.** 545 **REMARK.—If the vessel is quite full and closed on top, except at $O'$ where it communicates by a stationary pipe with a reservoir, the free surface cannot be formed, but the pressure at any point in the water is just the same during uniform rotation, as if a free surface were formed with vertex at $O$; i.e., $p_s = p_a + (h_s + z)g$. . . . (4) See figure for $h_s$ and $z$. (In subsequent paragraphs of this chapter the liquid will be at rest.) Fig. 474. **428a. Pressure on the Bottom of a Vessel containing Liquid at Rest.—If the bottom of the vessel is plane and horizontal, the intensity of pressure upon it is the same at all points, being** Fig. 475. Fig. 476. $p = p_a + h_y$ (Figs. 475 and 476), and the pressures on the elements of the surface form a set of parallel (vertical) forces. This is true even if the side of the vessel overhangs, Fig. 476, the resultant fluid pressure on the bottom in both cases being $$P = Fp - Fp_a = Fhy.$$ . . . . . . . . . . . . . . . . . . (1) (Atmospheric pressure is supposed to act under the bottom.) It is further evident that if the bottom is a rigid homogeneous plate and has no support at its edges, it may be supported at a 546 MECHANICS OF ENGINEERING. single point (Fig. 477), which in this case (horizontal plate) is its centre of gravity. This point is called the Centre of Pressure, or the point of application of the resultant of all the fluid pressures acting on the plate. The present case is such that these pressures reduce to a single resultant, but this is not always practicable. **Example.—In Fig. 476 (cylindrical vessel containing water), given $A = 20$ ft., $h = 15$ ft., $r_1 = 2$ ft., $r_2 = 3$ ft., find the pressure at the bottom, the vertical tension in the cylindrical wall $C_4$, and the hoop tension ($\tau$ 426) at $C$. (Ft., lb., sec.) Press on bottom = $F_{hy} = w r_1 h y = 16 \times 90 \times 69.5 = 62857$ lbs.; while the upward pull on $C_4 =$ $$\left(w r_1 - \pi r_1^2 h y\right) y = (16 - 1)\frac{90}{2} \times 69.5 = 33357 \text{ lbs.}$$ If the vertical wall is $t = \frac{1}{16}$ inch thick at $C$ this tension will be borne by a ring-shaped cross-section of area = $2 \pi r t$ (nearly) = $2 \pi \cdot 48 \times \frac{1}{16} = 30.17$ sq. inches, giving $(33357 + 30.17) = 1200$ lbs. per sq. inch shearing stress (vertical). The hoop tension at $C$ is horizontal and is $$p'' = r_1(p_p - p_o) + t$$ (see § 426), where $p_p = p_o + A_y;$$ $$\therefore p' = \frac{48 \times 15 \times 12}{\frac{1}{16}}(82.5 + 178) = 8125 \text{ lbs. per sq. in.}$$ (using the inch and poundy). **498. Centre of Pressure.—In subsequent work in this chapter, since the atmosphere has access both to the free surface of liquid and to the outside of the vessel walls, and $p_o$ would cancel out in finding the resultant fluid pressure on any elementary area $dF$ of those walls, we shall write: "The resultant fluid pressure on any dF of the vessel wall is normal to its surface and is $dp = pdF = sydF$, in which $s$ is the vertical distance of the element below the free surface of the liquid (i.e., $s$ is the "head of water"). If the surface pressed on is plane, these elementary pressures form a system of parallel forces, and may be replaced by a single resultant." A diagram showing a cylindrical vessel with a horizontal plate at the bottom, labeled "Centre of Pressure" and "Resultant Force". CENTRE OF PRESSURE. 547 (if the plate is rigid) which will equal their sum, and whose point of application, called the Centre of Pressure, may be located by the equations of § 23, put into calculus form. If the fluid be at rest, the resultant of a system of forces in space, and hence (§ 38) cannot in general be reduced to a single resultant, but to two, the point of application of one of which is arbitrary (viz., the arbitrary origin, § 38). Of course, the object of replacing a set of fluid pressures by a single resultant is for convenience in examining the equilibrium, or stability, of a rigid body the forces acting on which include those fluid pressures. As to their effect in distorting the rigid body, the fluid pressures must be considered in their true position (see example in § 364), and cannot be replaced by a resultant. 430. Resultant Liquid Pressure on a Plane Surface forming Part of a Vessel Wall. Co-ordinates of the Centre of Pressure—Fig. 478. Let $AB$ be a portion (of any shape) of a plane surface, horizontal, sustaining liquid pressure. Prolong the plane of $AB$ till it intersects the free surface of the liquid. Take this intersection as an axis $X$, $O$ being any point on $X$. The axis $X$, $\gamma$ to $Y$, lies in the given plane. Let $\alpha$ angle between $XY$ and the free surface. Then $x$ and $y$ are the co-ordinates of any elementary area $dF$ of the surface, referred to $X$ and $Y$. $z=$ the "head of water," below the free surface, of any $dF$. The pressures are parallel. The normal pressure on any $dF = ydzF$; hence the sum of these, is their resultant, $$P = \gamma y dz F = F_{xy} ; \ldots \ldots \ldots (1)$$ 548 MECHANICS OF ENGINEERING. in which $s$ is the "mean $z$" i.e., the $z$ of the centre of gravity $G$ of the plane figure $AB$, and $F^2$ total area of $AB (Fx = fdy)$, from eq. (4), § 23. $\gamma$ = heaviness of liquid (see § 400). That is, the total liquid pressure on a plane figure is equal to the weight of an imaginary prism of the liquid having a base equal to the area of the figure, and altitude equal to the depth of the centre of gravity of the figure below the surface of the liquid. For example, if the figure is a rectangle with one base (length = $b$) in the surface, and lying in a vertical plane, $$P = bh \cdot \gamma b y = \frac{bh}{2} \gamma y.$$ Evidently, if the altitude be increased, $P$ varies as its square. From (1) it is evident that the total pressure does not depend on the horizontal extent of the water in the reservoir. Now let $x_*$ and $y_*$ denote the co-ordinates, in plane $OXY$, of the centre of pressure, $C$, or point of application of the resultant pressure $P$, and apply the principle that the sum of the moments of each of several parallel forces, about an axis to them, is equal to the moment of their resultant about the same axis (§ 22). First taking $OF$ as an axis of moments, and then $OX$, we have $$Pz_* = \int_A (\text{syd} F) x, \quad \text{and} \quad Py_* = \int_A (\text{syd} F) y.$$ (3) But $Pz_* = Fz_*(\sin \alpha)y_*$, and the $s$ of any $dF = z dx$. Hence eq. (3) become (after cancelling the constant, $\gamma sin \alpha$) $$x_* = \frac{f^2 dF}{Fz} = \frac{I_y}{Fz}, \quad y_* = \frac{fy dy}{Fz}, \ldots$$ (3) in which $I_y$ is the "mom. of inertia" of the plane figure referred to $Y$ (see § 85). [N.B.—The centre of pressure thus found is identical with the centre of oscillation (§ 117) and the centre of percussion (§ 113) of a thin homogeneous plate, referred to axes $X$ and $Y$, $Y$ being the axis of suspension.] Evidently, if the plane figure is vertical $\alpha = 90^\circ$, $z = 0$ for CENTRE OF PRESSURE. 549 all $dF_a$ and $\overline{x} = \bar{x}$. It is also noteworthy that the position of the centre of pressure is independent of $\sigma$. **Norm.—Since the pressures on the equal $dF_a$ lying in any horizontal strip of the plane figure form a set of equal parallel forces equally spaced along the strip, and are therefore equivalent to their sum applied in the middle of the strip, it follows that for rectangles and triangles with horizontal bases, the centre of pressure must lie on the straight line on which the middles of all horizontal strips are situated.** **430. Centre of Pressure for Rectangles and Triangles with Bases Horizontal.—Since all the $dF_a$ of one horizontal strip have the same $\sigma$, we may take the area of this strip for $dF$ in the summation $dF/dx$. Hence for the rectangle $AB$, Fig 479, we have from eq. (8), § 430, with $dF = bdx$, $$x_a = K\frac{b}{h_1}\int_{h_1}^{h_2} x dx = \frac{2}{3}\frac{h_2 - h_1}{h_1^2}(1)$$ while (see note, § 430) $y_a = \frac{1}{2}b$. When the upper base lies in the surface, $h_1 = 0$, and $x_a = \frac{1}{2}h_2 = \frac{1}{2}$ of the altitude. For a triangle with its base horizontal and vertex up, Fig. 480, the length $u$ of a horizontal strip is variable and $dF = udx$. From similar triangles $u = \frac{h_2 - h_1}{h_1}(x - h_1)$; therefore $$x_a = \frac{b}{h_2 - h_1}\int_{h_1}^{h_2} (x - h_1) dx = \frac{b}{(h_2 - h_1)}(h_2 + \frac{1}{2}(h_2 - h_1))$$ But $\int_{h_1}^{h_2} (x - h_1) dx = \int_{-h_1}^{h_2 - h_1} x dx$ $$= \frac{1}{2}(3h_2 + h_1 - 4h_1)$$ $$= \frac{1}{2}(h_2 - h_1)(3a + 2ah + a^*)$$ $$= x_a = \frac{3a + 2ah + a^*}{2(h_2 + h_1)}$$ Fig. 479 Fig. 480 (9) 550 MECHANICS OF ENGINEERING. Also, since the centre of pressure must lie on the line $AB$ joining the vertex to the middle of base (see note, § 430), we easily determine its position. Evidently for $h_1 = 0$, i.e., when the vertex is in the surface, $$x_1 = \frac{1}{2} h_1.$$ Similarly, for a triangle with base horizontal and vertex down, Fig. 481, $$x = \frac{1}{2} (3h_1 + 2h_2 + h_3).$$ If the base is in the surface, $h_1 = 0$ and (3) reduces to $x_1 = \frac{1}{2} h_2$. It will be noticed that in this case the triangle the value of $x_1$ is the same whether be in shape so long as $h_1$ and $A$, remain unchanged and the base is horizontal. If the base is not horizontal, we may easily, by one horizontal line, divide the triangle into two triangles whose bases are horizontal and whose combined areas make up the area of the first. The resultant pressure on each of the component triangles is easily found by the foregoing principles, as also its point of application. The resultant of the two parallel forces so determined will act at some point on the line joining the centres of pressure of these two component pressures but not necessarily by the method of moments, while the amount of this final resultant pressure is the sum of its two components, since the latter are parallel. An instance of this procedure will be given in Example 3 of § 433. Similarly, the rectangle of Fig. 479 may be distorted into an oblique parallelogram with horizontal bases without affecting the value of $x_1$, nor the amount of resultant pressure, so long as $A$, and $h_1$ remain unchanged. 432. Centre of Pressure of Circle—Fig. 482. It will lie on the vertical diameter. Let $r$ = radius. From eq. (3), § 431, $$x_1 = \frac{I_r}{F r} = \frac{L_r + F^2 x_2}{F^2 x_2} = \frac{\pi r^2 - \pi r^2 x_2}{\pi r^2}.$$ (See eq. (4), § 88, and also § 91.) $$x_1 = x + \frac{1}{4} x^2.$$ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6) Fig. 481 Fig. 482 **CENTRE OF PRESSURE.** 551 **433. Examples.—It will be noticed that although the total pressure on the plane figure depends for its value upon the head, $z$, of the centre of gravity, its point of application is always lower than the centre of gravity. **EXAMPLE 1.—If 6 ft. of a vertical sluice-gate, 4 ft. wide, Fig. 483, is below the water-surface, the total water pressure against it is (ft. lb., sec.; eq. (1), § 430) $$P = F_{\text{avg}} = 6 \times 4 \times 3 \times 62.5 = 4500 \text{ lbs},$$ and (so far as the pressures on the vertical posts on which the gate slides are concerned) is equivalent to a single horizontal force of that value applied at a distance $x = \frac{6}{4} = 1.5$ ft. below the surface (§ 431). **EXAMPLE 2.—To begin by lifting the gate in Fig. 483, the gate itself weighing 900 lbs., and the coefficient of friction between the gate and posts being $\mu = 0.40$ (abstract number). (see § 156), we must employ an upward vertical force at least $$P' = 900 + 0.40 \times 4500 = 2000 \text{ lbs}.$$ **EXAMPLE 3.—It is required to find the resultant hydrostatic pressure on the trapezoid in Fig. 483a with the dimensions given and its bases horizontal; also its point of application, i.e., the centre of pressure of the plane figure in the position there shown. From symmetry the C. of P. will be in the middle vertical of the figure, at a distance $y$ from $AB$. It will be seen that of the two triangles $ABE$ and $CDF$ taken together (conceived to be shifted horizontally so that $CF$ and $BE$ coincide on the middle vertical), thus forming a single triangle of 5 ft. base, and having the same total pressure and C. of P. as the two actual triangles taken together. Let $P'$ be the total pressure, and $x'$ refer to the C. of P., for the rectangle; $P$, and $x$, for the 5 ft. tri-** A diagram showing a trapezoid with dimensions labeled A, B, C, D, E, F. Fig. 483a A diagram showing a trapezoid with dimensions labeled A, B, C, D, E, F. Fig. 483b 652 MECHANICS OF ENGINEERING. angle; $h_1 = 4$ ft. and $h_2 = 10$ ft. being the same for both. Then from eq. (1), § 480, we have (with the ft., lb., and sec.) $$P_1 = 30 \times 7y = 210y; \quad P_2 = \frac{1}{2} \times 6 \times 5 \times 6y = 90y;$$ while from eqs. (1) and (3) of § 481 we have also (respectively) $$z'_* = \frac{2}{3} \cdot \frac{1000 - 64}{84} = \frac{2}{3} \cdot \frac{936}{84} = 7.488 \text{ feet};$$ $$x'_* = \frac{1}{2} \cdot \frac{48 + 80 + 100}{8 + 10} = \frac{228}{2 \times 18} = 6.338 \text{ feet}.$$ The total pressure on the trapezoid, being the resultant of $P_1$ and $P_2$, has an amount $P = P_1 + P_2$ (since they are parallel), and has a lever arm $x$, about the axis $OY$ to be found by the principle of moments, as follows: $$\alpha_x' = \frac{P_1 x'_* + P_2 x'_*}{P_1 + P_2} = \frac{(210 \times 7.488 + 90 \times 6.338)y}{(210 + 90)y} = 7.09 \text{ ft.}$$ The total hydrostatic pressure on the trapezoid is (for fresh water) $$P = P_1 + P_2 = [210 + 90] \times 62.5 = 18750 \text{ lbs}.$$ **EXAMPLE 4** – Required the horizontal force $P'$, Fig. 484, to be applied at $N$ (with a leverage of $a = 30$ inches about the fulcrum $M$) necessary to (begin with) lift the circular disk $AB$ of radius $r = 5$ inches, covered by an opening of equal size. $NMAB$ is a single rigid lever weighing $G' = 210$ lbs. The centre of gravity, $G$, of disk, being a vertical distance $z = OG' = 40$ inches from the surface, is 50 inches (viz., the sum of $OM = k + r + z$ and $MG' = b - z$ from axis $OY$ is 90 inches). The centre of gravity of the whole lever is a horizontal distance $b' = 12$ inches, from $M$. Fig. 484 653 **EXAMPLES—CENTRE OF PRESSURE.** For impending lifting we must have, for equilibrium of the lever, $$P'a' = G'y + P(x_a - k); \quad . . . . (1)$$ where $P$ = total water pressure on circular disk, and $x_a = OC$. From eq. (1), § 430, (using inch, lb., and sec.) $$P = F_{y'} = \pi r^2 y' = \pi 100 \times 40 \times 62.5 = 454.6 \text{ lbs}.$$ From § 430, $z_c = OC = z + \frac{1}{4} \frac{r^2}{g} = 50 + \frac{1}{4} \frac{100}{50} = 50.5 \text{ in}.$ $$\therefore P = \frac{1}{30} [G'y + P(x_a - k)]$$ $$= \frac{1}{30} [210 \times 19 + 454.6 \times 30.5] = 546 \text{ lbs}.$$ **434. Example of Flood-gate.—Fig. 485.** Supposing the rigid double gate $AD$, 8 ft. in width, to have four hinges; two at $e$, and two at $f$, 1 ft. from top and bottom of water channel; required the pressures upon them, taking dimensions from the figure (ft., lb., sec.). Wat. press. = $P = F_{y'}$ $$= 78 \times 42 \times 62.5 = 20250$$ pounds, and its point of application (cent. of press.) is a distance $z_e = \frac{1}{2}$ of $9' = 6'$ from $O$ (§ 431). Considering the whole gate free and taking moments about $e$, we shall have (press. at $f$) $\times T = 20250 \times 5$; $\therefore$ press. at $f$ = 14464 lbs. (half on each hinge at $f$), and $\therefore$ press. at $e$ = $P$ - press. at $f$ = 5875 lbs. (half coming on each hinge). Fig. 485. 554 MECHANICS OF ENGINEERING. If the two gates do not form a single rigid body, and hence are not in the same plane when closed, a wedge-like or toggle-joint action is induced, producing much greater thrusts against the hinges, and each of these thrusts is not $\gamma$ to the plane of the corresponding gate. Such a case forms a good exercise for the student. 438. Stability of a Vertical Rectangular Wall against Water Pressure on One Side—Fig. 486. All dimensions are shown in the figure, except $l$, which is the length of wall $\gamma$ to paper. Supposing the wall to be a single rigid block, its weight $G^{\prime}$ = $h^{\prime} \gamma l^{\prime}$ ($\gamma$ being its heaviness (§ 7), and $l'$ its length). Given the water depth $h$, required the proper width $g$ for stability against tipping: First, the resultant of $G^{\prime}$ and the water-pressure $P$ must fall within the base $BD$ (or, which amounts to the same thing), the moment of $G^{\prime}$ about $D$, the outer toe of the wall, must be numerically greater than that of $P$; and Secondly, $P$ must be less than the sliding friction $f G^{\prime}$ (see § 166) on the base $BD$. Thirdly, the maximum pressure per unit of area on the base must not exceed a safe value (compare § 348). Now $P = F \frac{dy}{dx} = h \frac{dy}{dx} \gamma = \frac{1}{2} h^{\prime} \gamma l^{\prime} (\gamma = \text{heaviness of water})$; and $x_{c} = \frac{1}{2} l$. Hence for stability against tipping about $D$, $$P_{th} \text{ must be } < G^{\prime} f G^{\prime}; \text{i.e., } h^{\prime} \gamma l^{\prime} < \frac{1}{2} h^{\prime} \gamma l^{\prime}; \quad (1)$$ while, as to sliding on the base, $$P \text{ must be } < f G^{\prime}; \text{i.e., } h^{\prime} \gamma l^{\prime} < f b h^{\prime} l^{\prime}. \quad (2)$$ As for values of the coefficient of friction $f$, on the base of wall, Mr. Fanning quotes the following among others, from various authorities: Fig. 486. STABILITY OF WALL. 555 For point-dressed granite on dry clay, $f = 0.51$ moist clay, $0.33$ gravel, $0.58$ smooth concrete, $0.62$ similar granite, $0.70$ brickwork, $0.60$ For dressed hard limestone on like limestone, $0.38$ common bricks on common bricks, $0.64$ To satisfactorily investigate the third condition requires the detail of the next paragraph. 436. Parallelepipedal Reservoir Walls. More Detailed and Exact Solution.—If (1) in the last paragraph were an exact equality, instead of an inequality, the resultant $R$ of $P$ and $G$ would pass through the corner $D$, tipping would be impending, and thus the wall would be theoretically infinite. To avoid this we wish the wall to be wide enough that the resultant $R$, Fig. 487, may cut $BD$ in such a point, $E'$, as to cause the pressure per unit area, $p_n$, at $D$ to have a definite safe value (for the pressure $p_n$ at $D$, or quite near $D$, will evidently be greater than elsewhere on $BD$; i.e., it is the maximum pressure to be found on $BD$). This must be done by the principle of § 846 and can be done as follows: First, assume that $R$ cuts $BD$ outside of the middle third; i.e., that $$VE' = nb', > \frac{1}{2}b'(\text{or } n > 1);$$ where $n$ denotes the ratio of the distance of $E'$ from the middle of the base to the whole width, $b'$, of base. Then the pressure (per unit area) on small equal elements of the base $BD$ (see § 346) may be considered to vary as the ordinates of triangle $BDE'$ (i.e., as being within the distance $BD$), and $\overline{ED}$ will = $\frac{1}{2}MD$; i.e., A diagram showing a parallelepipedal reservoir wall with a section cut through it. The wall is labeled with points A, B, C, D, E', F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z. Points A and B are connected by line AB. Point C is connected to points D and E'. Line DE' intersects BD at point E'. Points F and G are connected by line FG. Points H and I are connected by line HI. Points J and K are connected by line JK. Points L and M are connected by line LM. Points N and O are connected by line NO. Points P and Q are connected by line PQ. Points R and S are connected by line RS. Points T and U are connected by line TU. Points V and W are connected by line VW. Points X and Y are connected by line XY. Points Z and A are connected by line AZ. 556 MECHANICS OF ENGINEERING. $$MD = 3(\frac{1}{n} - n)\psi.$$ The mean pressure per unit area, on $MD$, $$= G' + (L \cdot MD)_{h},$$ and hence the maximum pressure (viz., at $D$), being double the mean, is $$p_m = 2G' + [3\frac{1}{n}(\frac{1}{n} - n)] ; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (0)$$ and if $p_m$ is to equal $C'$ (see §§ 201 and 203), a safe value for the crushing resistance, per unit area, of the material, we shall have $$V(\frac{1}{n} - n)C' = 4G' = 4\frac{1}{n}h'\psi',$$ $$\therefore n = \frac{1}{2} - \frac{2}{3}\frac{h'}{\psi'} ; \quad . . . . . . . . . (1)$$ To find $\beta'$, knowing $n$, we put the $\Sigma$ (moms.) of the $G'$ and $P$ at $E'$, about $E'$, zero (for the only forces acting on the wall are the pressures of the foundation against it, along $MD$; and since the resultant of these latter passes through $E'$, the sum of their moments about $E'$ is already zero); i.e., $$G'n'b' - P'h = 0; \mathrm{~or~,~} nb'\frac{h'}{\psi'} = h\cdot\frac{1}{n}\psi';$$ $$\therefore V = h\sqrt{\frac{h'y'}{8nb'\psi'}} ; \quad . . . . . . (8)$$ Having obtained $\beta'$, we must also ascertain if $P$ is $ $\phi$, the angle of friction, with the vertical.) If $n$, computed from (1), should prove to be < $\frac{1}{4}$, our first assumption is wrong, and we therefore assume $n > \frac{1}{4}$, and proceed thus: Secondly, $n$ being < $\frac{1}{4}$ (see §§ 346 and 369), we have a STABILITY OF RESERVOIR WALLS. 557 trapezoid of pressures, instead of a triangle, on $BD$. Let the pressure per unit area at $D$ be $p_m$ (the maximum on base). The whole base now receives pressure, the mean pressure (per unit area) being $G' + \left(\frac{b}{2}\right)$; and therefore, from § 362, Case 1, we have $$p_m = [6n+1]\frac{G'}{b\ell} ; \quad \ldots \ldots \ldots (a)$$ and since, here, $G' = b'\kappa l'y'$, we may write $$p_a = (6n+1)\kappa l'y'$$ For safety as to crushing resistance we put $$6(n+1)\kappa y' = C'; \text{ whence } n = \frac{1}{6}\left[\frac{C'}{\kappa y'} - 1\right]. \quad (1a)$$ Having found $n$ from eq. (1a), we determine the proper width of base $b'$ from eq. (2), in case the assumption $s < \frac{1}{4}$ is verified. **EXAMPLE**—In Fig. 486, let $h' = 12$ ft., $h = 10$ ft., while the masonry weighs ($y'$) = 150 lbs. per cub. ft. Supposing it desirable to bring no greater compressive stress than 100 lbs. per sq. inch (= 14400 lbs. per sq. ft.) on the cement of the joints, we put $C' = 14400$, using the ft.-lb.-sec. system of units. Assuming $n > \frac{1}{4}$, we use eq. (1a) and obtain $$n = \frac{1}{2} \cdot \frac{2}{3} \cdot \frac{12 \times 150}{14400} = \frac{5}{12},$$ which is > $\frac{1}{4}$; hence the assumption is confirmed, also the propriety of using eq. (1) rather than (1a). Passing to eq. (2), we have $$b' = 10 \times \sqrt{\frac{62.5 \times 10^3}{\frac{1}{4} \times 12 \times 150}} = 3.7 \text{ feet}.$$ But, as regards frictional stability, we find that, with $f = 0.30$, a low value, and $b' = 3.7$ ft. (ft., lb., sec.), 658 MECHANICS OF ENGINEERING. $$\frac{P}{fG} = \frac{16V}{fVhLr'} = \frac{100 \times 62.5}{2 \times 0.3 \times 3.7 \times 12 \times 150} = 1.5;$$ which is greater than unity, showing the friction to be insufficient to prevent sliding (with $f=0.8$); a greater width must therefore be chosen, for frictional stability. If we make $n=\frac{1}{4}$, i.e., make $R$ cut the base at the outer edge of middle third ($\S$ 362), we have, from eq. (2), $$b' = 10 \times \sqrt{\frac{62.5 \times 10}{4 \times 12 \times 150}} = 5.89 \text{ feet};$$ and the pressure at $D$ is now of course well within the safe limit; while as regards friction we find $$P + fG' = 0.92 < \text{unity},$$ and therefore the wall is safe in this respect also. With a width of base = 3.7 feet first obtained, the portion $MD$, Fig. 457, of the base which receives pressure (according to Navier's theory) is about one sixth of the base in length, or about one sixth of the base, the portion $BM$ tending to open, and perhaps actually suffering tension, if capable (i.e., if cemented to a rock foundation), in which case these tensions should properly be taken into account, as with beams ($\S$ 295), thus modifying the results. It has been considered safe by some designers of high masonry dams, to neglect these possible tensile resistances, as it has just been done in deriving $\theta = 3.7$ feet; but others, in view of the great importance of this factor of safety in the character of Navier's theory, when applied to the very wide bases of such structures, prefer, in using the theory (as the best available), to keep the resultant pressure within the middle third at the base (and also at all horizontal beds above the base), and thus avoid the chances of tensile stresses. This latter plan is supported by Messrs. Church and Felsy, as engineers of the proposed Quaker Bridge Dam in connection with the New Croton Aqueduct of New York City, in their report of 1867. See $\S$ 363. A diagram illustrating mechanics of engineering. RESERVOIR WALLS. 559 437. Wall of Trapezoidal Profile. Water-face Vertical. Economy of material is favored by using a trapezoidal profile, Fig. 488. With this form the stability may be investigated in a corresponding manner. The portion of wall above each horizontal bed should be examined similarly. The weight $G'$ acts through the centre of gravity of the whole mass. Details of Fig. 488 show the vertical cross-section of a trapezoidal wall, with notation for dimensions as indicated; the portion considered having a length = $l$, $\gamma$ to the paper. Let $\gamma$ = heaviness of water, $\gamma'$ that of the masonry (assumed homogeneous), with $n$ as in § 436. For a triangle of pressure, $MD$, on the base, i.e., with $n > \frac{1}{2}$, or resultant falling outside the middle third (neglecting possibility of tensile stresses on left of $M$), if the intensity of pressure $p_m$ at $D$ is to be $C'$ (§ 305), we put, as in § 436, $$b^2(1-n)C' = \frac{1}{2}G', \text{i.e., } = \frac{1}{2}h'\cdot(b'+b'')\gamma',$$ whence $$n = \frac{1}{2} - \frac{1}{3}\frac{b'}{b''} - \frac{b'}{b}, \quad \ldots \quad (1)$$ For a trapezoid of pressure, i.e. with $n < \frac{1}{2}$, or the resultant of $P$ and $G'$ falling within the middle third, we have, as before (§ 362, Case I), $$p_m,\text{or } C' = (6n + 1)\frac{G'}{bL};$$ whence $$n = \frac{1}{6}\left[\frac{9CL}{G'} - 1\right]; \text{i.e., } n = \frac{1}{6}\left[\frac{9CL}{bL\gamma'(b+b'')} - 1\right]. \quad (1a)$$ 560 MECHANICS OF ENGINEERING. From the geometry of the figure, having joined the middles of the two bases, we have $$\bar{y}_i = g \bar{O} = \frac{h'}{3} b' + 2b''$$ (§ 26, Prob. 6), and, by similar triangles, $\overline{OV}: \overline{KP}:: g \bar{O}: h'$, whence $$\overline{OV} = \frac{\bar{y}}{h'}, \frac{1}{2}(b' - b'')$$ :. $\overline{OE_1} = \overline{OF} + \overline{VF'} = \frac{1}{6}\left(\frac{b' + 2b''}{b'}\right) + nb'. . . (a)$$ The lines of action of $G'$ and $P$ meet at $E$, and their resultant cuts the base in some point $E'$. The sum of their moments about $E'$ should be zero, i.e., $F \cdot A = G' \cdot OE_1$ that is, (see eq. (a) above, and eq. (1), § 430), $$M_y = l h' y' \frac{b'}{b'} + b'' \gamma_1 \left[\frac{(b' + 2b'')(b' - b')}{b'} + nb'\right] . . . (b)$$ i.e., cancelling, $$M_y = l h' y'\left[(b' + 2b'')(b' - b') + 6nb'(b' + b'')\right]. . . (8')$$ Hence we have two equations for finding two unknowns viz: (1)' and (2)' when $n > \frac{1}{4}$; and (1a)' and (2)' when $n < \frac{1}{4}$. For dams of small height (less than 40 ft., say), if we immediately put $n = \frac{1}{4}$, thus restricting the resultant pressure to the edge of the middle third, and solve (2)' for $y$, we get, assuming a somewhat probable value of $h'$ for a footwalk, or roadway, while $h'$ may be taken enough greater than $h$ to provide against the greatest height of waves, from 2.5 to 6 ft., the value of $p_w$ at $D$ will probably be < $C'$. In any case, for a value of $n =$, or < $\frac{1}{4}$ we put $p_w$ for $C'$ in equation (1a)' and solve for $p_w$, to determine if it is no greater than $C'$. Mr. Fanning recommends the following values for $C'$ (in lbs.)
$n$$C'$
$\frac{1}{4}$300 lbs.
$\frac{1}{5}$270 lbs.
$\frac{1}{6}$250 lbs.
$\frac{1}{8}$230 lbs.
560 RESEVOIR WALLS. 561 per sq. foot) with coursed rubble masonry laid in strong mortar:
For Limestone. Sandstone. Granite. Brick.
$C'$ 50,000 50,000 60,000 35,000
Average heaviness of the masonry in lbs. per cub. ft. 152 132 154 120
As to frictional resistance, $P$ must be $ \frac{1}{4} \quad \ldots \quad (1)''$$ and $$p_m = \frac{h'y}{6n+1} \text{ for } n < \frac{1}{4} \quad \ldots \quad (1a)''$$ ($p_m$ not to exceed $C'$ in any case); while to determine the breadth of base, $b'$, after $s$ is computed [or assumed, for small height of wall], we have from eq. (2), $$h'y = \frac{h'b'y}{6n+1} \quad \ldots \quad (3)''$$ Also, for frictional stability, $$\frac{h'y}{b'} \text{ must be } < f\frac{h'b'y}{6n+1} \quad \ldots \quad (8)''$$ 562 MECHANICS OF ENGINEERING. 438. High Masonry Dams.—Although the principle of the arch may be utilized for vertical stone dikes of small height (30 to 50 feet) and small spans for greater heights and spans the formula for hoop tension, § 426 (or rather, here, “hoop compression”), on the vertical radial joints of the horizontal arch rings, Fig. 489, calls for so great a radial thickness of joint in the lower course, that a straight dike would be “gravity dam” and usually built instead, even where firm rock abutments are available laterally. For example, at a depth of 100 feet, where the hydrostatic pressure is $A p = 100 \times 62.5 = 6250$ lbs. per sq. ft., if we assume for the voussoirs a (radial, horizontal) thickness $= 4$ ft., with a (horizontal) radius of curvature $r = 100$ feet, we shall find a compression between their vertical radial faces of (ft. lb. sec.) $$p' = \frac{r^2(p - p_a)}{\ell} = \frac{100 \times 6250}{4} = 156250 \text{ lbs. per sq. ft.},$$ or 1085 lbs. per sq. inch; far too great for safety, even if there were no danger of collapse, the dike being short. If now the thickness is increased, in order to distribute the pressure over a greater surface, we are met by the fact that the formula for “hoop compression” is no longer strictly applicable, the law of distribution of pressure becoming very uncertain; and even supposing that uniformity of stress can be maintained, the thickness demanded for proper safety against crushing is greater than for a straight dam (“gravity dam”) at a very moderate depth below the water surface, unless the radius of curvature of arch can be made small. But the smaller the radius the more does the dam encroach on the storage capacity of the reservoir, while in no case, of course, can it be made smaller than half the span. Another point is, that as masonry is not destitute of elasticity, the longer the span the more unlikely is it that the parts of the arch will “close up” properly, and develop its A diagram showing a cross-section of a high masonry dam. Fig. 489 HIGH MASONRY DAMS. 563 abutment reactions when the water is first admitted to the reservoir; which should occur if it is to act as an arch instead of by gravity resistance. For many years the engineers of the proposed Quaker Bridge Dam reported unfavorably to the plan of a curved design for that structure, and recommended that a straight dam be built. See reference in § 486. According to their designs this dam is to be 258 feet in height (which exceeds by about 90 feet the height of any dam previously built), about 1400 feet in length at the top, and 216 feet in width at the lowest point of base, joining the bed-rock. More recently, however (1888), a board of experts, specially appointed for that purpose, have examined a number of different plans, have reported favorably to the adoption of a curved form for the dam, as offering greater resistance under extraordinary circumstances (impact of ice-floes, earthquakes, etc.), on account of its arched form (though resisting by gravity action under usual conditions) than a straight structure; and also as more pleasing in appearance. Fig. 490 shows the profile of a straight high masonry dam as designed at the present day. Assuming a width $b'$ = from 6 to 29 feet, and a height $h$ = from 100 to 200 feet, and exceeding the maximum height of waves, the upstream outline $ACM$ is made nearly vertical and perhaps somewhat concave, while the down-stream profile $BDN$, by computation or graphical trial, or both, is so formed that when the reservoir is full the resultant $R$, of the weight $$G$$ of the portion $ABCD$ of masonry above each horizontal bed, as $CD$, and the hydrostatic pressure on the downstream face $AC$, shall cut the bed $OD$ in such a point $E'$ as not to cause too great compression $\rho_m$ at the outer edge $D$ (not over 5 lbs. per sq. inch according to M. Krantz in "Reservoir Waller"), $\rho_m$ being computed by one of the equations (10) and (6a) of § 436) A diagram showing the profile of a straight high masonry dam with an arched upstream outline ACM and a concave down-stream profile BDN. Fig. 490 564 MECHANICS OF ENGINEERING. For $E'$ outside the middle third $\frac{1}{3} \cdot p_m = \frac{2G}{3CD(1-n)}$ (1)** and neglecting tension For $E'$ inside middle third $\frac{1}{3} \cdot p_m = \frac{(6n+1)G}{CD}$ (1e)** where $l =$ length of wall $\gamma$ to paper, usually taken = one foot, or one inch, according to the unit of length adopted; for $n$, see § 436. Nor, when the reservoir is empty and the water pressure lacking, must the weight $Q$ resting on each bed, as $CD$, cut the bed in a point $E'$ so near the edge $C$ as to produce excessive pressure there (computed as above). The figure shows the general form of profile resulting from these conditions. The masonry may be laid in any manner, by irregular bedding in every direction as to make the wall if possible a monolith. For more detail see next paragraph. 440. Quaker Bridge Dam (on the New Croton Aqueduct). Attempts, by strict analysis, to determine the equation of the curve $BN$, $AM$ being assumed straight, so as to bring the point $E'$ on the outer edge of the middle third at its joint, or nearly so, the result is that $D$ lies below a definite line, have failed, up to the present time; but approximate and tentative methods are in use which serve all practical purposes. As an illustration the method set forth in the report on the Quaker Bridge Dam will be briefly outlined; this method confines $E'$ to the middle third. The width $AB = b'$ is taken = 2 ft for a roadway, and $h' = \gamma$ ft. The profile is made a vertical rectangle from $A$ down to a depth of 1 ft., representing water surface (reservoir), (fig. 70). Considering the weight of this rectangular body with the corresponding water pressure (for a length of wall = one foot), we find the resultant pressure comes a little within the outer edge of the middle third of the base of the rectangle, while $p_m$ is of course small. The rectangular form of profile might be continued below this horizontal joint, as far as complying with the mode QUAKER BRIDGE DAM. 565 third requirement, and the limitation of pressure-intensity, is concerned; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make the widening of the joints too abrupt; but, not to make Having computed (by means) this trapezoid and constructed its diagram (fig. 3), by drawing a great number of trapezoids, the value of the resultant $G$ of this weight and that of the rectangle is found (by principle of moments or by an equilibrium polygon) in amount and position and combined with water pressure corresponding $44$ ft. width. The force $R$, whose point intersection with new joint or bed (lower base trapezoid) is noted and value $p_m$ computed. These should both be somewhat nearer their limits than in preceding joint. If $E'$ different width should be chosen for trapezoidal layers at this factory. Similarly another layer $11$ ft. height and trapezoidal form is added below and treated in same way and so on until joint at depth $66$ ft. from water surface a width is found where point $E'$ being close upon its limiting position while $p_m$ quite a little under limit set for upper joint dam $8$ tons per square foot. For next three $11$ ft. trapezoidal layers chief governing element middle third requirement $E'$ being kept quite close limit while increase $p_m$ 7.85 tons per square foot possible when point $E''$ left-hand edge left vertical so that reservoir empty point $E'''$ shall be near bottom upstream edge $C$. Down depth about 200 ft. value $p_m$ allowed increase 10.48 tons per sq. ft., while position $E'$ gradually retreats from edge its limit Beyond 300 ft. depth prevent rapid increase width consequent extreme flattening flow downstream course $p_m$ allowed maintain constant 7.85 tons per square foot which value reaches at point $N$ base 366 MECHANICS OF ENGINEERING. the dam, which has a width = 216 ft., and is 258 feet below the water surface when the reservoir is full. The heaviness of the masonry is taken as $y' = 156.25$ lbs. per cubic foot, just $\frac{1}{4}$ of $y = 62.5$ lbs. per cub. foot, the heavi- ness taken for water. When the reservoir is empty, we have the weight $G$ of the superincumbent mass resting on any bed $CD$, and applied through the point $E'$ ; the pressure per unit area at $C$ can therefore be determined by means of § 80, a better method of $\left(\frac{1}{4}CD + CE'\right) + CD$ for this purpose. In the present case we find $E'$ to be within middle third at all joints, and the pressures at $C$ to be under the limit. For further details the reader is referred to the report itself (reprinted in Engineering News, January, 1888, p. 20). The graphic results were checked by computation, Wegmann's method, applied to each trapezoid in turn. 441. Earthwork Dam, of Trapezoidal Section.—Fig. 401. It is required to find the conditions of sta- bility of the straight earthwork dam $ABDE$, whose length $l_1$ L to paper, as regards sliding horizontally on the plane $AE$; i.e., its frictional stability. With the dimensions of the figure, $y$ and $y'$ being the heavi- nesses of the water and earth respec- tively (see § 7), we have Weight of dam = $G = \text{vol} \times y' = b_1[l + h(a - o)]y'$. (1) Resultant water press. = $P = F \cdot y = OA \times l \times hy'$. (2) Horiz. comp. of $P = H = P \sin \alpha$ = $(OA \sin \alpha)hy' = hhy'$. . . . (3) From (3) it is evident that the horizontal component of $P$ is just the same, viz., $= hl$. Hy, as the water pressure would be on a vertical rectangle equal to the vertical projection of $OA$ A diagram showing a trapezoidal section of an earthwork dam with labeled points A, B, C, D, E, and lines indicating various measurements. Fig. 401 EARTHWORK DAM. 567 and with its centre of gravity at the same depth (§4). Compare § 416. Also, $$Vert.\ comp.\ of\ P=V=Pcos\alpha$$ $$=(OAd\ cos\alpha)\#dly=\#dly,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (4)$$ and in the same as the water pressure on the horizontal projection of $OA$ if placed at a depth = $OG = hA$. For stability against sliding, the horizontal component of $P$ must be less than the friction due to the total vertical pressure on the plane $AE$, viz., $G + V$; hence if $\gamma$ is the coefficient of friction on $AE$, we must have $H < f(G + V)$, i.e. (see above), $$\#dly must be < f[\theta,(b+(a+c))y+\#dly]\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .(5)$$ However, if the water leak under the dam on the surface $AE$, so as to exert an upward hydrostatic pressure $$V'=[a+b+c]h_{y},$$ (to make an extreme supposition,) the friction will be only $$=f(G+V-V'),$$ and (8) will be replaced by $$H < f(G+V-V').$$ Experiment shows (Weisbach) that $\gamma = 0.38$, computations made from (6) (treated as a bare equality) give satisfactory results. EXAMPLE.-(Ft., lb., sec.) With $\gamma = 0.38$, $h = 20$ ft., $a = 22$ ft., $b = 24$ ft., $c = 26.4$ ft., and $z = 30$ ft., we have, making (6) an equality, with $\gamma' = 3y$, $$#dly=f[yth,(b+\frac{a+c}{2})+\#dly-(a+b+c)\#dly]\ ,$$ :. (400) = #[(32(0+28.2)(0+24×20)-(32×20)+(32×20)+b+h)]\(30);\) whence, solving for $b$, the width of top, $b = 10.3$ feet. 668 MECHANICS OF ENGINEERING. 449. Liquid Pressure on Both Sides of a Gate or Rigid Plate. The sluice-gate $AB$, for example, Fig. 492, receives a pressure, $P_1$, from the "head-water" $M$, and an opposing pressure $P_2$ from the "tail-water" $N$. Since these two horizontal forces are not in the same line, though parallel, their resultant $R$, which = $P_1 - P_2$, acts horizontally in the same plane, but at a distance below $O_1 = u$, which we may find by placing the moment of $R$ about $O_1$, equal to the algebraic sum of those of $F_1$ and $F_2$, about $O_1$, \[ \therefore \quad Ru = P_1 x'_1 - P_2 (x'_2 + h) \quad \ldots \quad (1) \] \[ \therefore \quad u = \frac{[P_1 x'_1 - P_2 (x'_2 + h)]}{P_1 - P_2} \quad \ldots \quad (2) \] $C_1$ and $C_2$ are the respective centres of pressure of the surfaces $O_1B$ and $O_2B$, and $u =$ distance of $R$ from $O_1$, while $h =$ difference of level between head and tail waters. If the surfaces $O_1B$ and $O_2B$ are both rectangular, $x'_1 = 4A,$ and $x'_2 = 4B.$ EXAMPLE.—Let the dimensions be as in Fig. 493, both surfaces under pressure being rectangular and 8 ft. wide. Then (ft., lb., sec.), $R = F' - F''$, (cf. § 430) \[ R = [12 \times 8 \times 6 - 8 \times 8 \times 4]62.5 \] \[ = 9000 \text{ lbs.} = 10 \text{ tons}; \] while from ex. (2) \[ u = \frac{[12 \times 8 \times 6 \times 8 - 8 \times 8 \times 4(4)]62.5}{9000}. \] That is, $u = 6.93$ feet, which locates $C$. Hence the pressure of the gate upon its hinges or other support is the same (aside WATER ON BOTH SIDES OF GATE. 569 from its own weight), provided it is rigid, as if the single horizontal force $R = 10$ tons acted at the point $C$, 2.93 ft. below the level of the tail-water surface. 443. If the plate, or gate, is entirely below the tail-water surface, the resultant pressure is applied in the centre of gravity of the plate.—Proof as follows: Conceive the surface to be divided into a great number of small equal areas, each $dF$; then, the head of water of any $dF$ being $x$, on the head-water side, and $y$ on the tail-water side, the resultant pressure on $dF$ is $ydx dy dF$, in which $A$ is the difference of level between head and tail water. That is, the resultant pressures on the equal $dF$'s are equal, and hence form a system of equal parallel forces distributed over the plate in the same manner as the weights of the corresponding portions of the plate; therefore their single resultant acts through the centre of gravity of the plate; Q.E.D. This single resultant $= ydx dy dF = y dx dy F = Fh_y$ EXAMPLES: Fig. 87. The resultant pressure on a circular disk of radius $r = 8$ inches, acting through the vertical partition $O K$, which has its centre of gravity 3 ft. below the tail-water surface, with $h = 2$ ft., is (ft., lb., sec.) $$R = Fh_y = w r h_y$$ $$= \pi^2 \times 24 \times 2.93 = 174.6 \text{ lbs}.$$ (and is applied through the centre of gravity of the circle.) Evidently $R$ is the same for any depth below the tail-water surface, so long as $h = 2$ ft. [Let the student find a graphic proof of this statement.] 444. Liquid Pressure on Curved Surfaces.—If the rigid surface is curved, the pressures on the individual $dF$'s, or elements of area, do not form a system of parallel forces, and the single resultant (if one is obtainable) is not equal to their sum. In A diagram showing a curved plate with a vertical line indicating its center of gravity at a height \( h \) from the tail-water surface. The plate is divided into small areas \( dF \), each with a head \( x \) above and \( y \) below the center of gravity. The pressure on each area is shown as \( ydx dy dF \). The total pressure \( R \) is calculated as \( Fh_y \), where \( F \) is the weight per unit area and \( h_y \) is the vertical distance from the center of gravity to the tail-water surface.] 570 MECHANICS OF ENGINEERING. general, the system is not equivalent to a single force, but can always be reduced to two forces (§ 38) the point of application of one of which is arbitrary (the arbitrary origin of $S$) and its amount is $\nu(2XY + (2YZ) + (2Z))$. A single Example will be given of a thin rigid shell having a circular base, the surface of a right cone, Fig. 495, its altitude being $h$ and radius of base $r$. It has no bottom, is placed on a smooth horizontal table, vertex up, and is filled with water through a small hole in the apex $O$, which is left open (to admit atmospheric pressure). What load, besides its own weight $G$, must be placed upon it to prevent the water from lifting it and escaping under the table? The total force acting on each $dF$ of the inner curved surface is $sydF$ and is normal to the surface. Its vertical compon. is $sydF \cos \alpha$, and horizontal compon. = $sydF \cos \alpha$. The $dF's$ have all the same $\alpha$, but different $z's$ (or heads of water). The lifting tendency of the water on the thin shell is due to the vertical components forming a system of || forces, while the horizontal components, radiating symmetrically from the axis of the cone, neutralize each other. Hence the resultant lifting force is $$V = Z(\text{vert. comp.)} = y\sin\alpha \cdot sdF = y\sin\alpha \cdot Fz; \quad (1)$$ where $F =$ total area of curved surface, and $\bar{z}$ = the "head of water" of its centre of gravity. Eq. (1) may also be written thus: $$V = yF_{b}\bar{z}; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)$$ in which $F_{b} = F\sin\alpha =$ area of the circular base = area of the projection of the curved surface upon a plane $\gamma$ to the vertical, i.e., upon a horizontal plane. Hence we may write $$V = \frac{4}{3}\pi r^2 h; \quad . . . . . . . . . . . . . . . . . . . . . . (3)$$ since $\bar{z} = \frac{4}{3}h$, being the $z$ of the centre of gravity of the curved CONICAL SHELL. 571 surface and not that of the base. $y = \text{heaviness of water}$. If $G' = \text{weight of the shell}$ and is $< V$, an additional load of $V - G'$ will be needed to prevent the lifting. If the shell has a bottom of weight $= G''$, forming a base for the cone and rigidly attached to it, we find that the vertical forces acting on the whole rigid body, base and all, are: $V$ upward; $G'$ and $G''$ downward; and the liquid pressure on the base, viz., $V = \pi r^2 h y$ (§ 430) also downward. Hence the resultant vertical force to be counteracted by the table is downward, and $$= G' + G'' + V' - V,$$ which = $G' + G'' + \frac{\pi r^2 h y}{\pi r^2 h y}; \quad (4)$$ i.e., the total weight of the rigid vessel and the water in it, as we know, of course, in advance. CHAPTER III. EARTH PRESSURE AND RETAINING WALLS [Note.—This chapter was outlined and written mainly by Prof. C. L. Crandall, and is here incorporated with his permission. The theory of earth pressure is arranged from Bau- meister.] 445. Angle of Repose.—Granular materials, like dry sand, loose earth, soil, gravel, pease, shot, etc., on account of the friction between the component grains, occupy an intermediate position between liquids and large rigid bodies. When heaped up, they will stand at an angle of inclination with the horizontal greater than a definite angle called the **angle of natural slope**, or **angle of repose**, different for each material; so that if the side of the mass is to be retained permanently at some greater angle, a **Retaining Wall** (or "Revetment Wall," in military parlance) becomes necessary to support it. If the material is somewhat moist it may be made to stand alone at an inclination greater than that of the natural slope, on account of the cohesion thus produced, but only as long as this degree of moisture exists. If no cohesion whatever is present, it assumes the consistency of mud and may require a much thicker wall, if it is to be supported laterally, than if dry. In dealing with earth to be supported by a retaining wall, we consider the former to have lost any original cohesion which may have existed among its particles, or that it will eventually lose it through the action of the weather; and hence treat it as a granular material. A few approximate values of the angle of natural slope are 573 RETAINING WALLS. 573 given below, being taken from Fanning, p. 345; see reference on p. 588 of this work.
MATERIAL Angle of Repose Coefficient of Friction Ratio of Slope
Dry sand, fine. 36° 0.88 Horsi. to vert.
" coarse. 36° 0.77 1.78" 1
Damp clay. 42° 1.000 1.00" 1
Wet clay. 42° 0.95 1.05" 1
Clayey gravel. 45° 1.000 1.00" 1
Sand. 45° 0.95 1.05" 1
Gravel. 36° 0.781 1.38" 1
Firm soil. 36° 0.781 1.38" 1
Vegetable soil. 36° 0.700 1.43" 1
Peat. 36° 0.644 2.75" 1
The angle of repose, or natural slope, is also, evidently, the angle of friction between two masses of the same granular material. 446. Earth Pressure, and Wedge of Maximum Thrust.—Fig. 496. Let $AB$ be a retaining wall, having a plane face $AB$ in contact with a mass of earth $ABD$, both wall and earth being of indefinite extent $\gamma$ to the paper. Let $AD$ be the natural slope of the earth, making an angle $\beta$ with the vertical (see preceding table), and let $AC$ be the angle of repose; see preceding table). Since $AB$, making an angle $\alpha$ with the vertical, is more nearly vertical than $AD$, the retaining wall is necessary, to keep the mass $ABD$ in the position shown. The profile $BCD$ may be of any form in this general discussion. Suppose the wall to be on the point of giving way; then the following motions are impending: 1st. Sliding is impending between some portion $ABC'A'$ of the mass of earth and the remainder $C'AD$, the surface of rupture $AC'D$ being in figure, but not forming a plane, but lying somewhere on the profile $(BCD)$ being assumed plane, and making some angle $\phi'$ (to be determined) with the vertical. At this instant the resultant pressure $N'$ of $AC'D$ on the plane $AC'$ of the mass $ABC'$ (a wedge) must make an angle $\beta = \text{comp. of angle of friction with } AC'$ on the upper side. A diagram showing a retaining wall (AB) in contact with a mass of earth (ABD), with angles of repose and friction indicated. 2nd. Rotation about point A is impending, so that the plane $ABD$ will become horizontal at some instant after sliding has commenced. 3rd. Rotation about point B is impending, so that the plane $ABD$ will become vertical at some instant after sliding has commenced. 4th. Rotation about point C is impending, so that the plane $ABCD$ will become horizontal at some instant after sliding has commenced. 5th. Rotation about point D is impending, so that the plane $ABCD$ will become vertical at some instant after sliding has commenced. 6th. Rotation about point E is impending, so that the plane $EBCD$ will become horizontal at some instant after sliding has commenced. 7th. Rotation about point F is impending, so that the plane $EFCD$ will become vertical at some instant after sliding has commenced. 8th. Rotation about point G is impending, so that the plane $GBCD$ will become horizontal at some instant after sliding has commenced. 9th. Rotation about point H is impending, so that the plane $HBCD$ will become vertical at some instant after sliding has commenced. 10th. Rotation about point I is impending, so that the plane $IBC'D$ will become horizontal at some instant after sliding has commenced. 11th. Rotation about point J is impending, so that the plane $JBC'D$ will become vertical at some instant after sliding has commenced. 12th. Rotation about point K is impending, so that the plane $KBC'D$ will become horizontal at some instant after sliding has commenced. 13th. Rotation about point L is impending, so that the plane $LBC'D$ will become vertical at some instant after sliding has commenced. 14th. Rotation about point M is impending, so that the plane $MBC'D$ will become horizontal at some instant after sliding has commenced. 15th. Rotation about point N is impending, so that the plane $NBC'D$ will become vertical at some instant after sliding has commenced. 16th. Rotation about point O is impending, so that the plane $OBC'D$ will become horizontal at some instant after sliding has commenced. 17th. Rotation about point P is impending, so that the plane $PBC'D$ will become vertical at some instant after sliding has commenced. 18th. Rotation about point Q is impending, so that the plane $QBC'D$ will become horizontal at some instant after sliding has commenced. 19th. Rotation about point R is impending, so that the plane $RBC'D$ will become vertical at some instant after sliding has commenced. 20th. Rotation about point S is impending, so that the plane $SCD'$ will become horizontal at some instant after sliding has commenced. 21st. Rotation about point T is impending, so that the plane $TCD'$ will become vertical at some instant after sliding has commenced. 22nd. Rotation about point U is impending, so that the plane $UCD'$ will become horizontal at some instant after sliding has commenced. 23rd. Rotation about point V is impending, so that the plane $VCD'$ will become vertical at some instant after sliding has commenced. 24th. Rotation about point W is impending, so that the plane $WCD'$ will become horizontal at some instant after sliding has commenced. 25th. Rotation about point X is impending, so that the plane $XCD'$ will become vertical at some instant after sliding has commenced. 26th. Rotation about point Y is impending, so that the plane $YCD'$ will become horizontal at some instant after sliding has commenced. 27th. Rotation about point Z is impending, so that the plane $ZCD'$ will become vertical at some instant after sliding has commenced. 28th. Rotation about point A' is impending, so that the plane $A'B'C'D'$ will become horizontal at some instant after sliding has commenced. 29th. Rotation about point B' is impending, so that the plane $B'C'D'$ will become vertical at some instant after sliding has commenced. 30th. Rotation about point C' is impending, so that the plane $C'D'$ will become horizontal at some instant after sliding has commenced. 31st. Rotation about point D' is impending, so that the plane $D'$ will become vertical at some instant after sliding has commenced. 32nd. Rotation about point A'' is impending, so that the plane $A''B''C''D''$ will become horizontal at some instant after sliding has commenced. 33rd. Rotation about point B'' is impending, so that the plane $B''C''D''$ will become vertical at some instant after sliding has commenced. 34th. Rotation about point C'' is impending, so that the plane $C''D''$ will become horizontal at some instant after sliding has commenced. 35th. Rotation about point D'' is impending, so that the plane $D''$ will become vertical at some instant after sliding has commenced. 36th. Rotation about point A''' is impending, so that the plane $A'''B'''C'''D''' will be horizontal at some instant after sliding has commenced. 496 37th. Rotation about point B''' is impending, so that the plane B'''C'''D''' will be vertical at some instant after sliding has commenced. 496 38th. Rotation about point C''' is impending, so that the plane C'''D''' will be horizontal at some instant after sliding has commenced. 496 39th. Rotation about point D''' is impending, so that the plane D''' will be vertical at some instant after sliding has commenced. 496 40th. Rotation about point A'''' is impending, so that the plane A''''B''''C''''D'''' will be horizontal at some instant after sliding has commenced. 496 41st. Rotation about point B'''' is impending, so that the plane B''''C''''D'''' will be vertical at some instant after sliding has commenced. 496 42nd. Rotation about point C'''' is impending, so that the plane C''''D'''' will be horizontal at some instant after sliding has commenced. 496 43rd. Rotation about point D'''' is impending, so that the plane D'''' will be vertical at some instant after sliding has commenced. 496 44th. Rotation about point A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D'''', A'''', B'''', C'''', D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D ''' ,A ''' ,B ''' ,C ''' ,D '''. 574 MECHANICS OF ENGINEERING. 2d. A downward sliding of the mass $AB'C$ along the back face $AB$ of the wall. That is, the resultant pressure $P'$ of the wall against the mass $BAC$ at this instant makes an angle $\theta$ (= complement of angle of friction between the earth and wall) with the plane $AB$ and on the upper side. The weight of the wedge of earth $BAC'$ will be called $G'$, and we desire to find the pressure $P'$ against the wall. Let $BAC$ be a wedge (of the earth-mass), in which $AC$ makes any angle $\phi$ with $A V_1$, and suppose it to be on the point of moving down and forcing out the wall; thus encountering friction both on the plane $AC$ and the plane $AB$. Then the forces acting on it are, viz.: $G$, its own weight, vertical; $N$, the resultant pressure of the earth below it, making an angle $\beta$ with $AC$ on upper side; and $P$, the resultant pressure of the wall, at angle $\theta$ with $AB$ (see Fig. 496 for positions of $N$ and $P$). If now we express the force $P'$ in terms of $\phi$ and other quantities, and find that value $\phi'$, of $\phi$, for which $P'$ is a maximum, we thereby determine the "wedge of maximum thrust," $ABC'A$; while this maximum thrust, $F'$, is the force which the wall must be designed to resist so that the wall is overturned, the earth will sink with it until this part of its surface gradually assumes the natural slope.] Let $G =$ weight of prism of base $ABC$, and altitude = unity $\gamma$ to paper; then $G = \gamma \times area ABC$, where $\gamma =$ "heaviness" = wgt. per cub. unit, of earth. Now $P$, $G$, and $\delta$ balance; therefore, in triangle abc, if ab and ac are drawn || A diagram showing a wedge-shaped block (BAC) resting on a wall (AB). The wedge makes an angle φ with AC on its upper side. The forces acting on the wedge are labeled: G (weight), N (resultant pressure of earth below), P' (resultant pressure of wall), and θ (angle between P' and AB). The diagram also shows a line segment from N to P', representing the maximum thrust F'. The text explains that by finding the value φ' for which P' is a maximum, one determines the "wedge of maximum thrust," ABC'A. Fig. 496 RETAINING WALLS. 575 and $G$ and $N$ respectively, $\delta c$ is = and $\parallel$ to $P$; and from Trigonometry we have $$P=G\frac{\sin(\beta-\phi)}{\sin(\beta+\delta-\phi)};\quad\quad(1)$$ in which $\delta$ stands for $\alpha+\theta$, for brevity, being the angle which $P$ makes with the vertical. $N$ makes an angle = $\beta-\phi$ with the vertical. The value, $\phi'$, of $\phi$, which makes $P$ a maximum is found by placing $\frac{dP}{d\phi}=0$. From eq. (1), remembering that $G$ is a function of $\phi$, and that $\beta$ and $\delta$ are constants, we have $$\frac{dP}{d\phi}=\frac{\sin(\beta+\delta-\phi)\left[\frac{dG}{d\phi}\sin(\beta-\phi)-G\cos(\beta-\phi)+G\sin(\beta-\phi)\cos(\beta+\delta-\phi)\right]}{\sin(\beta+\delta-\phi)}$$ For $P$ to be a maximum we must put numerator of above = 0. . . . . . . . . . . . . . . . . . . . . . . . . . (a) To find a geometrical equivalent of $\frac{dG}{d\phi}$ denote $AU$ by $L$, and draw $AE$, making an angle = $\delta\phi$ with $AC$. Now the area $ACI=AI\times\frac{1}{2}CE=(L+dL)\times Ld\phi=\frac{1}{2}dL^2\ldots$ (neglecting infinitesimal of 3rd order). Now $$dG=y\times\text{area }ACI\times\text{unity};:\frac{dG}{d\phi}=\frac{dy}{dx}L^2;\therefore(a) \text{ becomes}$$ $$\sin(\beta+\delta-\phi)dL^2\sin(\beta-\phi)-\sin(\beta+\delta-\phi)G\cos(\beta-\phi)$$ $$+G\sin(\beta-\phi)\cos(\beta+\delta-\phi)=0;$$ i.e., $G=$ $$\frac{\frac{dy}{dx}L^2\sin(\beta-\phi)\sin(\beta+\delta-\phi)}{\sin(\beta+\delta-\phi)\cos(\beta-\phi)-\cos(\beta+\delta-\phi)\sin(\beta-\phi)}$$ 576 MECHANICS OF ENGINEERING. when $P$ is a maximum; and hence, calling $G'$ and $\phi'$ and $L'$ the values of $G$, $\phi$, and $L$, for max. $P$, we have $$G' = \frac{1}{2} y L' \sin(\beta - \delta) \sin(\beta + \delta - \phi'), \quad (3)$$ and therefore from (1) $P$ max. itself is $$P' = \frac{1}{2} y L' \cdot \frac{\sin(\beta - \phi')}{\sin \delta}, \quad (3)$$ 447. Geometric Interpretation and Construction.--If in Fig. 496 we draw $CF$, making angle $\delta$ with $AD$, $C$ being any point on the ground surface $BD$, we have $$CF = L' \cdot \frac{\sin(\beta - \phi)}{\sin \delta}$$ Drop a perpendicular $FH$ from $F$ to $AC$, and we shall have $$FH = CF' \cdot \sin(\beta + \delta - \phi') = L' \cdot \frac{\sin(\beta - \phi) \sin(\beta + \delta - \phi)}{\sin \delta}$$ From this it follows that the weight of prism of base $ACF$ and unit height $$= y L'. FH = y L' \cdot \frac{\sin(\beta - \phi) \sin(\beta + \delta - \phi)}{\sin \delta} \quad (4)$$ When $AC$ (as $\phi$ varies) assumes the position and value $AC'$ bounding the prism of maximum thrust, Fig. 497, $L$ becomes $y L'$ and $\phi = \phi'$; and eq. (4) gives the weight of the prism $AC'F'$ when its section is secant to one of the prism (or wedge) of maximum thrust $ABC'$, by comparing eq. (4) with eq. (3); that is, $AC'$ bisects the area $ABCF'$, and hence may be determined by fixing such a point $C'$ on the upper profile $BD$, as to make the triangular area $ACF'$ equal to the sectional area of the wedge $BC'A'C'F'$ being drawn at an angle $\delta$ with $AD$. This holds for any form of ground surface $BD$, or as RETAINING WALLS. 577 values of the constants $\beta$, $\alpha$, or $\theta$. $C'$ is best found graphically by trial, in dealing with an irregular profile $BD$. Having found $AC' = L'$, $P'$ can be found from (3), or may readily be fol-lows: (Fig. 487) With $F'$ as a centre and radius = $C'F'$, describe an arc cutting $AD$ in $J'$, and join $C'J'$. The weight of prism with base $C'J'F'$ and unit height will = $P'$. For that prism has a weight $$= 4Y \cdot F'J' \cdot C'H';$$ but $$F'J' = F'C' = \frac{L'\sin(\beta - \phi)}{\sin\delta},$$ and $$CH' = L'\sin(\beta - \phi);$$ :. weight of prism $C'J'F$ = $4Y L'\sin(\beta - \phi)/\sin\delta$ = $P'$. [See eq. (3).] 448. Point of Application of the Present Earth Thrust.--This thrust (called $P'$ throughout this chapter except in the present paragraph) is now known in magnitude and direction, but not in position; i.e., we must still determine its line of action, etc. Divide $AB$ into a number of equal parts, $ab$, $bc$, $cd$, etc.; see Fig. 488. Treat $ab$ as a small retaining wall, and find the magnitude $P'$ of the thrust against it by § 447; treat $ac$ similarly, thus finding the thrust, $P''$, against it; then $ad$, $ae$, etc., the thrusts against them being found to be $P'''$, $P''r$, etc.; and so on. Now the pressure $$P'$$ on ab is applied nearly at middle of ab, $$P'' - P'''$$ " " " " " " bc, $$P'''' - P'''$$ " " " " " cd, A diagram showing a retaining wall with a cut-out section labeled AB. Points A, B, C, D are marked on the wall. Lines AC', AD', and AH are drawn from points A, D, and H respectively to points C', J', and H'. The diagram also shows lines C'H', CH', and F'C'. The text describes the calculation of the weight of a prism with base C'J'F' and unit height. Fig. 487 578 MECHANICS OF ENGINEERING. and so on. Erect perpendiculars at the middle points of $ab$, $bc$, $cd$, etc., equal respectively to $P'$, $P''$, $P'''$, etc., and join the ends of the perpendicular through the centre of gravity of the area so formed (Fig. 498), will give, on $AB$, the required point of application of the thrust or earth pressure on $AB$, and this, with the direction and magnitude already found in § 447, will completely determine the thrust against the wall $AB$. 449. Special Law of Loading.--If the material to be retained consists of loose stone, masses of masonry, buildings, or even moving loads, as in the case of a wharf or roadway, each can be replaced by the same weight of earth or other material which will render the bank homogeneous, situated on the same vertical, and the profile thus reduced can be treated by §§ 447 and 448. Should the solid mass extend below the plane of rupture, $AC$, and the plane of natural slope, it will become retaining wall for the material below it if strong enough to act upon such (limber) profile ($BCD$). Fig. 498 to the front of the mass, or to the front and line of rupture for maximum thrust above it, if it does not reach the surface; if not strong enough, or if it does not reach below the plane of natural slope, its presence is better ignored, probably, except that the increased weight must be considered. The spandrel wall of an arch may present two of these special cases; i.e., the profile may be enlarged to include a moving load, while it may be limited at the back by the other spandrels. If the earth profile starts at the front edge of the top of wall, instead of from the back as at $B$, Fig. 496, eq. (3) would only apply to the portion behind $AB$ prolonged, leaving the part on the wall (top) to be treated as a part of the wall to aid in resisting the thrust. If the wall is stepped in from the footings, or foundation Fig. 498. RETAINING WALLS. 579 courses, probably the weak section will be just above them; if stepped at intervals up the back of the wall, the surface of separation between the wall and filling, if it is plane, will probably pass through the first step and incline forward as much as possible without cutting the wall. **450. Straight Earth-profile.—The general case can be simplified as follows (the earth-profile $BD$ being straight, at angle $\zeta$ with vertical, = $DET$): Since the triangles $ABC'$ and** A diagram showing a straight earth profile with angles and lines labeled. **Fig. 68.** $CAF'$ are equal, from § 447, and $AC'$ is common, therefore $BS = F'H$ (both being drawn to $AC'$). Draw $AE$ and $BM \parallel$ to $F'C'$ (i.e., at angle $\phi$ with $AD$), cutting $DB$, prolonged, in $E$. We have $$\frac{DE}{CE} = \frac{EI}{EA - CE^2} \quad \text{and} \quad \frac{CE}{BE} = \frac{EA}{EA - BM'}$$ But $C'F' = BM$ (since $BS = H'F'$); therefore $$\frac{DE}{CE} = \frac{CE}{BE}; \text{i.e., } DE \cdot BE = CE^2,$$ which justifies the following construction for locating the desired point $C$ on $BD$, and thus finding $AC' = L'$ and the angle $\phi$: Describe a circle on $ED$ as a diameter, and draw 580 MECHANICS OF ENGINEERING. $BX \parallel BD$, thus fixing $X$ in the curve. With centre $E$ describe a circular arc through $X$, cutting $BD$ in $C'$, required. Having $AC' \parallel (i.e., L')$, $\phi'$ is known; hence from eq. (8) we obtain the earth thrust or pressure $P'$; or, with $P'$ as centre and radius $= C'C'$, describe a circle $C''$; then the triangle $C''F''K''$ is the base of a prism of unity height whose weight is $P''$ (as in § 467). **Centres of Pressure.—Applying the method of § 448, Fig. 498, to this case, we find that the successive $L''$'s are proportional to the depths $ab$, $ac$, $ad$, etc., and that the successive $P''$'s are proportional [see (6)] to the squares of the depths; hence the area in Fig. 498 must be triangular in this case, and the point of application of the resultant pressure on $AB$ is one third of $AB$ from $A$: just as with liquid pressure. **461. Resistance of Retaining Walls—(Fig. 500).** Knowing the height of the wall we can find its weight = $G'$, for an assumed thickness, and unity width $\gamma$ to paper. The resultant of $G'$, acting through the centre of gravity of wall, and $P'$, the thrust of the embankment, in its proper line of action, should cut the base $AV$ within the middle third and make an angle with the normal (to the base) less than half angle between walls. For the straight wall and straight earth-profile of Fig. 499 and § 450, the length $L' = AC'$, can be expressed in terms of the (vertical) height, $h$, of wall, thus: $$\overline{AB} = \frac{h}{\cos \alpha},$$ and $L' = AC' = AB \cdot \frac{\sin(\zeta - \alpha)}{\sin(\zeta - \phi')} = \frac{h}{\cos \alpha} \cdot \frac{\sin(\zeta - \alpha)}{\sin(\zeta - \phi')};$ ∴ eq. (8) becomes $$P' = h \cdot \frac{\sin(\beta - \phi') \cdot \sin(\zeta - \alpha)}{\sin \alpha} = h \cdot A.$$ (9) [A representing the large fraction for brevity.] RETAINING WALLS. 581 This equation will require, for a wall of rectangular section, that the thickness, $d$, increase as $h$, in order that its weight may increase as $A^{\prime}$ (i.e., as $P$) and that its resisting moment may increase with the overturning moment. By this equality of moments is meant that $Pa = Gb$; where $a$ and $b$ are the respective lever-arms of the two forces about the front edge of the middle third. ($AB$ is the back of the wall). In other words, their resultant will pass through this point. The following table is computed on the basis just mentioned, viz., that the resultant of $P'$ and $G$ shall pass through the front edge of the middle third. The symbols of eq. (5) and the table are all shown in Fig. 499, except $\theta$, $\varphi$, and $d$. $y =$ weight of a cubic foot of earth, here assumed = $\frac{1}{3}$ that of masonry (e.g., if earth weighs 100 lbs., masonry is assumed to weigh 150 lbs. per cubic foot); $\delta =$ angle between $PA$ and $AD$; $P' =$ force acting on the wall; and $\delta = \alpha + \theta$; $\delta =$ # in this case as the wall is vertical, or $\alpha = 0$. $d =$ the proper safe thickness to be given to the wall, of rectangular section, to prevent overturning, as stated above; $h =$ the altitude, and $A$ is the fraction shown in eq. (5). Whether the wall is safe against sliding on its base, and whether a safe compression per unit area is exceeded on the front edge of the base, are matters for separate consideration. The latter will seldom govern with ordinary retaining walls.
$\alpha = 0$; i.e., wall is vertical; also density of wall = $\frac{1}{3}$ that of the earth. L II III
$\frac{c}{2} = \frac{27}{2}$ $\frac{c}{2} = \frac{27}{2}$ $\frac{c}{2} = \frac{27}{2}$
tan $\beta$ $\theta$ $d$ $d'$ $d'$ $d'$ $d'$ $d'$
1.0 45° 291°.17 364 28°.18 364 45° 71°.30h
1.5 36° 289°.66 364 28°.18 364 45° 71°.30h
2.0 33° 287°.12 364 28°.18 364 45° 71°.30h
4.0%
In Case I of table, since $\alpha = 0$, $\theta = 90°$ and $\zeta = 90°$; $\therefore \delta = 90°$, and hence $C'F'$ of Fig. 499 is $\gamma$ to $AD$, so that 583 MECHANICS OF ENGINEERING. (since the area of $\triangle ABC' = \triangle AC'F'$) $\phi'$ must = $\beta$. These values, in (5), give $$P' = \frac{1}{2}yA'\tan\beta\beta; \text{i.e., } A = \tan\beta\beta.\ . . . \ (6)$$ In Case II, since $\zeta = 90^\circ$, $\alpha = 0$ and $\theta = \beta$, i.e., $\delta = \beta$; and (6) reduces to $$P' = \frac{1}{2}yA'\sin(\beta - \phi'); \text{i.e., } A = \frac{\sin(\beta - \phi')}{\sin\beta\cos\phi'}.\ . . . \ (7)$$ In Case III, $\zeta = \beta$ and $BD$ will be || to $AD$, $D$ being at infinity. See Fig. 501. Through $B$ draw $BH$ ⊥ to $AD$, and $BF'$ making angle $\delta$ with $AD$, $C'$ is now to be located on $BD$, so as to make (area of) $\triangle ABC' =$ (area of) $\triangle AC'F'$, according to (4). But $\angle C'AF' = C'FA$ being $\pi/4 + 44^\circ$, and $AC'F'A$ being $\delta = \alpha + \theta$; $\delta = 0$, in this case, and hence also = $\beta$. Conceive $B$ and $F'$ to be joined. Now $\triangle AC'F' = \triangle ABF'' = \triangle BF'F''$. But $\triangle ABC' = \triangle BF'F''$ (equal bases and altitudes). Hence $\triangle ABC'$ cannot = $\triangle AC'F'$ unless $C'$ is moved out to infinity; and then $\phi'$ becomes $\beta$, and eq. (5) reduces to $$P' = \frac{1}{2}yA'\sin\beta; \text{i.e., } A = \sin\beta.\ . . . \ (8)$$ [Increasing $\alpha$ from zero will decrease the thickness $d$: i.e., inclining the wall inwards will decrease the required thickness, but diminish the frictional stability at the base, unless the latter be ⊥ to $AB$. The back of the wall is frequently inclined outwards, making the section a trapezoid, to increase the frictional stability at the base when necessary, as with timber walls supporting water.] RETAINING WALLS. 583 **452. Practical Considerations.— An examination of the values of $A$ and $d$ in the table of p. 461 will show that in sup- porting quicksand and many kinds of clay which are almost fluid under the influence of water, it is important to know what kind of drainage can be secured, for on that will depend the thickness of the wall. With such conditions, free from water-bearing strata, an assumed natural slope of $1\frac{1}{4}$ to 1 (i.e., 14 hor. to 1 vert.), will be safe ; the actual pressure below the effect of frost and surface water will be that due to a much steeper slope on account of cohesion (neglected in this theory). The thrust from freely placed material can be reduced by depositing it in layers sloping back from the wall. If it is not so placed, however, the natural slope will seldom be flatter than 14 to 1 unless reduced by water. In supporting material which is liable to water-bearing at low slopes, such as the and overlain by strata which are liable to become semi-fluid and slippery, the thrust may exceed that due to semi-fluid ma- terial on account of the surcharge. If these strata are under the wall and cannot be reached by the foundation, or if resist- ance to sliding cannot be obtained from the material in front by sheet-piling, no amount of masonry can give security. Water at the back of the wall will, by freezing, cause the material to exert an indefinitely greater pressure, besides disinte- grating the soil. If this is dangerous, its accumulation, drainage should be provided by a layer of loose stone at the back leading to "weep-holes" through the wall. A friction-angle at the back of the wall equal to that of the filling should always be realized by making the back rough by steps, or projecting stones or bricks. Its effect on the required thickness is too great to be economically ignored. The resistance to slippage at the base can be increased, when necessary, by inclining the foundation inwards; by stepping or sloping back from the wall; by increasing its effective weight or incline the thrust more nearly to the vertical; by sheet-piling in front of the foundation, thus gaining the resist- ance offered by the piles to lateral motion; by deeper founda- tions, gaining the resistance of the earth in front of the wall.
452. Practical Considerations. An examination of the values of $A$ and $d$ in the table of p. 461 will show that in sup- porting quicksand and many kinds of clay which are almost fluid under the influence of water, it is important to know what kind of drainage can be secured, for on that will depend the thickness of the wall. With such conditions, free from water-bearing strata, an assumed natural slope of $1\frac{1}{4}$ to 1 (i.e., 14 hor. to 1 vert.), will be safe ; the actual pressure below the effect of frost and surface water will be that due to a much steeper slope on account of cohesion (neglected in this theory). The thrust from freely placed material can be reduced by depositing it in layers sloping back from the wall. If it is not so placed, however, the natural slope will seldom be flatter than 14 to 1 unless reduced by water. In supporting material which is liable to water-bearing at low slopes, such as the and overlain by strata which are liable to become semi-fluid and slippery, the thrust may exceed that due to semi-fluid ma- terial on account of the surcharge. If these strata are under the wall and cannot be reached by the foundation, or if resist- ance to sliding cannot be obtained from the material in front by sheet-piling, no amount of masonry can give security. Water at the back of the wall will, by freezing, cause the material to exert an indefinitely greater pressure, besides disinte- grating the soil. If this is dangerous, its accumulation, drainage should be provided by a layer of loose stone at the back leading to "weep-holes" through the wall. A friction-angle at the back of the wall equal to that of the filling should always be realized by making the back rough by steps, or projecting stones or bricks. Its effect on the required thickness is too great to be economically ignored. The resistance to slippage at the base can be increased, when necessary, by inclining the foundation inwards; by stepping or sloping back from the wall; by increasing its effective weight or incline the thrust more nearly to the vertical; by sheet-piling in front of the foundation, thus gaining the resist- ance offered by the piles to lateral motion; by deeper founda- tions, gaining the resistance of the earth in front of the wall.
584 MECHANICS OF ENGINEERING. The coefficient of friction on the base ranges, according to Trautwine, from 0.30 to 0.30 on wet clay; " .50 to .66 " dry earth ; " .60 to .75 " sand or gravel ; " .60 on a dry wooden platform ; to .75 on a wet one. If the wall is partially submerged, the buoyant effort should be subtracted from $G$, the weight of wall. 483. Results of Experience.--(Trautwine.) In railroad practice, a vertical wall of rectangular section, sustaining sand, gravel, or earth, level with the top [p. 682 of Civ. Eng. Pocket Book] and loosely deposited, as when dumped from cars, cars, etc., should have a thickness $d$, as follows:
If of cut stone, or of first-class large ranged rubble, in mortar. $d$ = .50
" good common scabbled mortar-rubble, or brick." $d$ = .45
" well scabbled dry rubble." $d$ = .50
Where $A$ includes the total height, or about 3 ft. of foundations. (a) For the best masonry of its class $h$ may be taken from the top of the foundation in front. (b) A mixture of sand or earth, with a large proportion of round boulders or cobbles, will weigh more than the backing assumed above; requiring $h$ to be increased from one eighth to one sixth of the height. (c) The wall will be stronger by inclining the back inwards, especially if of dry masonry, or if the backing is put in place before the mortar has set. (d) The back of the wall should be left rough to increase friction. (e) Where deep freezing occurs, the back should slope outwards for at least four feet below the top and be left smooth. (f) When a wall is too thin, it will generally fail by bulging outward at about one third the height. The failure is usually gradual and may take years. (g) Counterforts, or buttresses at the back of the wall, usually of rectangular section, may be regarded as a waste of masonry, although considerably used in Europe; the bond will RETAINING WALLS. 585 seidom hold them to the wall. Buttresses in front add to the strength, but are not common, on account of expense. (a) Land-ties of iron or wood, tying the wall to anchors imbedded below the line of natural slope, are sometimes used to increase stability. (i) Walls with curved cross-sections are not recommended. **454. Conclusions of Mr. B. Baker.—(Actual Lateral Pressure of Earthwork.)** Experience has shown that $d = 0.25$, with batter of 1 to 2 inches per foot on face, is sufficient when tacking and foundation are both favorable; also that under no ordinary conditions of arching or heavy backing, with solid foundation, it is necessary for $d$ to be greater than 0.60. Mr. Baker found that run-off water, running off the top of the footings, with a batter of 1½ inches per foot, in ground of average character; and, if any material is taken out to form a face-panel, three fourths of it is put back in the form of a plaster. The object of the batter, and of the panel if used, is to distribute the pressure better on the foundation. All the walls of the "District Railway" (London) were designed on this basis, and there has not been a single instance of settlement, or overturning, or of sliding forward. **455. Experiments with Models—Accounts of experiments with apparatus on a small scale, with sand, etc., may be found in vol. **LXXI** Proceedings of Institution of Civil Engineers, London: England (p. 350); also in vol. II of the "Annales des Ponts et Chaussées" for 1883 (p. 788). The following table shows the results and the results of experience given in §§ 453 and 454, when compared with the table of p. 581 indicate a fairly close agreement between practice and theory. This agreement is believed to be close enough so that the general method of §§ 447 and 451, with the table of p. 581, can be relied upon in practice. The greatest value of this method will, of course, be for cases of exceptional loading, inclined wall, etc., where the results of experience do not furnish so valuable a guide.
CHAPTER IV. HYDROSTATICS (Continued)--IMMERSION AND FLOTATION. 486. Rigid Body Immersed in a Liquid. Buoyant Effort.--If any portion of a body of homogeneous liquid at rest be conceived to become rigid without alteration of shape or bulk, it would evidently still remain at rest; i.e., its weight, applied at its centre of gravity, would be balanced by the pressures, on its bounding surfaces, of the contiguous portions of the liquid; hence, If a rigid body or solid is immersed in a liquid, both being at rest, the resultant action upon it of the surrounding liquid (or general situation) is called the "buoyant effort," equal in amount to the weight of liquid displaced, and acting through the centre of gravity of the volume (considered as homogeneous) of displacement (none occupied by the solid). This point is called the centre of buoyancy, and is sometimes spoken of as the centre of gravity of the displaced water. If $V$ = the volume of displacement, and $\gamma$ = heavi- ness of the liquid, then the $$\text{buoyant effort} = V\gamma.$$ (1) (By "volume of displacement" is meant, of course, the volume of liquid actually displaced when the body is immersed.) If the weight of the solid is not equal to the buoyant effort, or if the centre of gravity does not lie along a vertical as the centre of buoyancy, the two forces form an unbalanced system and motion begins. But as a consequence of this very motion the action of the liquid is modified in a manner dependent on the shape and kind of motion of the body. 588 **IMMERSION.** 587 Problems in this chapter are restricted to cases of rest, i.e., balanced forces. Suppose $G' = V'\gamma$; then, If the centre of gravity lies in the same vertical line as the centre of buoyancy and underneath the latter, the equilibrium is **stable**; i.e., after a slight angular disturbance the body returns to its original position (after several oscillations if above the equilibrium) and is unstable. If they coincide, as when the solid is homogeneous (but not hollow), and of the same heaviness ($\S 7$) as the liquid, the equilibrium is **indifferent**, i.e., possible in any position of the body. The following is interesting in this connection: In an account of the new British submarine boat "Nautilus," a writer in *Chambers's Journal* remarked (1887): "At each side of the vessel are four port-holes, into which fit cylinders two feet in diameter. When these are filled with air and turned outwards, they can be raised by a gear, but the displacement of the boat is so much increased that the vessel rises to the surface; but when the cylinders are withdrawn into their sockets, it will sink." As another case in point, large water-tight canvas "air-bags" have recently been used for raising sunken ships. They are sunk in a collapsed state, attached by divers to the submerged vessel, and then inflated with air from pumps above, which of course largely segments their displacement while adding no appreciable weight. **457. Examples of Immersion—Fig. 502.** At (a) is an example of stable equilibrium, the centre of buoyancy $B$ being above the centre of gravity $C$, and the buoyant effect $V'\gamma > G'$ being greater than the weight of the solid; at (b'), conversely, we have unstable equilibrium, with $V'\gamma$ still = $G'$. At (b) the buoyant effect $V'\gamma$ is > $G'$, and Fig. 502.; 588 MECHANICS OF ENGINEERING. to preserve equilibrium the body is attached by a cord to the bottom of the vessel. The tension in this cord is $$S_b = V'y - G' \qquad \qquad (1)$$ At $(c)$ $V'y$ is $< G'$, and the cord must be attached to a support above, and its tension is $$S_c = G' - V'y \qquad \qquad (9)$$ If in eq. (9) [(c) in figure] we call $S_c$ the apparent weight of the immersed body, and measure it by a spring- or beam-balance, we may write The apparent weight of a solid totally immersed in a liquid equals its real weight diminished by that of the amount of liquid displaced; in other words, the loss of weight = the weight of displaced liquid. **EXAMPLE 1.**—How great a mass (not hollow) of cast-iron can be supported in water by a wrought-iron cylinder weighing 140 lbs., if the latter contains a vacuous space and displaces 3 cubic feet of water, both bodies being completely immersed [Pt., lb., sec.] The buoyant effort on the cylinder is $$V'y = 3 \times 62.5 = 187.5 \text{ lbs},$$ leaving a residue of 47.5 lbs. upward force to buoy the cast-iron, whose volume $V'$ is unknown, while its heaviness $(8\ 7)$ is $y' = 450$ lbs. per cub. foot. The direct buoyant effort of the water on the cast-iron is $V''y' = [V' \times 62.5]$ lbs., and the problem requires that this force + 47.5 lbs. shall = $V''y''$ = the weight $G''$ of the cast-iron; :. $V'' \times 62.5 + 47.5 = V'' \times 450;$ :. $V'' = 0.19$ cub. ft., while $0.12 \times 450 = 54$ lbs. of cast-iron. **Ans.** **EXAMPLE 2.**—Required the volume $V'$, and heaviness $y'$, of a homogeneous solid which weighs 6 lbs. out of water and 4 lbs. when immersed (apparent weight) (ft., lb., sec.). A diagram showing a cylinder with a hole in it, partially submerged in water. IMMERSION. 689 From eq. (3), $4 = 6 - V' \times 62.5$; . . . $V' = 0.082$ cub. feet. :. $y' = G' + V' = 6 + 0.039 = 187.5$ lbs. per cub. ft., and the ratio of $y'$ to $y$ is $187.5 : 0.039 = 8.0$ (abstract number); i.e., the substance of this solid is three times as dense, or three times as heavy, as water. [The buoyant effort of the air has been neglected in giving the true weight at 6 lbs.] 458. Specific Gravity.—By specific gravity is meant the ratio of the heaviness of a given homogeneous substance to that of a standard homogeneous substance; in other words, the ratio of the weight of a certain volume of the substance to the weight of an equal volume of the standard substance. Distilled water at the temperature of maximum density (4° Centigrade) under a pressure of 14.7 lbs. per sq. inch is sometimes taken as the standard substance, more frequently, however, as sea-water. For any homogeneous substance, whatever its specific gravity with respect to the standard substance, the numerical example just given illustrates a common method of determining experimentally the specific gravity of a homogeneous solid substance, the value there obtained being 3. The symbol $\sigma$ will be used to denote specific gravity, which is evidently an abstract number. The standard substance should always be mentioned, and its heaviness $y$; then the heaviness of a substance whose specific gravity is $\sigma$ $$y' = \sigma y.$$ (1) and the weight $G'$ of any volume $V'$ of the substance may be written $$G' = V'y' = V'\sigma y.$$ (2) Evidently a knowledge of the value of $y'$ dispenses with the use of $\sigma$, though when the latter can be introduced into problems involving the buoyant effort of a liquid the criterion as to whether a homogeneous solid will sink or rise, when immersed in the standard liquid, is more easily applied, thus: Being immersed, the volume $V'$ of the body is that, $V_1$, of displaced liquid. Hence, A diagram showing a submerged object and displaced liquid. 590 MECHANICS OF ENGINEERING. if $G'$ is > $V'y$, i.e., if $V'y'$ is > $V'y$, or $\sigma > 1$, it sinks; while if $G'$ is < $V'y$, . . . . . . or $\sigma < 1$, it rises; i.e., according as the weight $G'$ is > or < than the buoyant effort. Other methods of determining the specific gravity of solids, liquids, and gases are given in works on Physics. 459. Equilibrium of Flotation.—In case the weight $G'$ of an immersed solid is less than the buoyant effort $V'y$ (where $V'$ is the volume of displacement, and $y$ the deviation of liquid) the body rises to the surface, and after a series of oscillations comes to rest in such a position, Fig. 560, that its centre of gravity $C$ and the centre of buoyancy $B$ (the new $B$, belonging to the new volume of displacement, which is limited above by the horizontal plane of the free surface of the liquid) are in the same vertical (called the axis of flotation, or line of support), and that the volume of displacement has diminished to such a new value $V'$, that $$V'y = G'. \quad \ldots \quad (1)$$ In the figure, $V =$ vol. AND, below the horizontal plane AN, and the slightest motion of the body will change the form of this volume, in general (whereas with complete immersion the volume of displacement remains constant). For stable equilibrium it is not essential in every case that $C$ (centre of gravity of body) coincides with $B$ (centre of buoyancy) as with complete immersion, since if the solid is turned, $B$ may change its position in the body, as the form of the volume AND changes. There is now no definite relation between the volume of displacement $V$ and that of the body, $V'$, unless the latter is homogeneous, and then for $G'$ we may write $V'y'$, i.e. $$V'y' = V'y \text{ (for a homogeneous solid)}; \quad \ldots \quad (2)$$ or, the volumes are inversely proportional to the heavinesses. Fig. 560 FLOTATION. 591 The buoyant effort of the air on the portion $ANE$ may be neglected in most practical cases, as being insignificant. If the solid is hollow, the position of its centre of gravity $C$ may be easily varied (by shifting ballast, e.g.) within certain limits, but that of the centre of buoyancy $B$ depends only on the geometrical form of the volume of displacement $\Delta V D$, below the water level. **Example.--(Ft., ib., sec.)** Will a solid weighing $G' = 400$ lbs., and having a volume $V' = 8$ cub. feet, without hollows or recesses, float in water? To obtain a buoyant effort of 400 lbs., we need a volume of displacement, see eq. (1), of $$V = \frac{G'}{\gamma} = \frac{400}{62.5} = \text{only } 6.4 \text{ cub. ft.}$$ Hence the solid will float with $8 - 6.4$, or 1.6, cub. ft. projecting above the water level. **Query:** A vessel contains water, reaching to its brim, and also a piece of ice which floats without touching the vessel. When the ice melts with the water overflow? **460. The Hydrometer is a floating instrument for determining the relative heaviness of liquids. Fig. 504 shows a simple form, consisting of a bulb and a cylindrical stem of glass, so designed and weighted as to float perfectly in all liquids whose heaviness is to be determined. Let $F$ denote the uniform sectional area of the stem (a circle), and suppose that when floating in water (whose heaviness = $\gamma$) the water surface marks a point $A$ on the stem; and that when floating in another liquid, say petroleum, whose heaviness = $\gamma_p$, we wish to determine, it floats at a greater depth, the liquid surface marking $A'$ on the stem; let height above $A$ be $G$. $G$ is the same in both experiments; but while the volume of displacement in water is $V$, in petroleum it is $V + Fz$. Therefore from eq. (1), § 459, Fig. 504. Fig. 504 592 MECHANICS OF ENGINEERING. in the water $G' = V\gamma$, . . . . . (1) and in the petroleum $G' = (V + Fz)\gamma_s$; . . . . (2) from which, knowing $G'$, $F$, $z$, and $\gamma$, we find $V$ and $\gamma_s$, i.e., $$V = \frac{G'}{\gamma} \text{ and } \gamma_s = \frac{G'\gamma}{G' + Fz\gamma}. \text{ . . . (3)}$$ [N.B. $-F$ is best determined by noting the additional distance, $= l$, through which the instrument sinks in water under an additional load $P$, not immersed; for then] $$G' + P = (V + Fd\gamma, \text{ or } F = \frac{P}{l\gamma}).$$ EXAMPLE.—[Using the inch, ounce, and second, in which system $\gamma = 1000 + 128 = 0.578 (\S 409).]$ With $G' = 3$ ounces, and $F = 0.10$ eq. then, $x$ being observed on the graduated stem, to be 5 inches, we have for the petroleum $$\gamma_s = \frac{3 \times 0.578}{3 + 0.10 \times 5 \times 0.578} = 0.525 \text{ oz. per cubic inch}$$ $$= 56.7 \text{ lbs. per cub. foot}.$$ Temperature influences the heaviness of most liquids to some extent. In another kind of instrument a scale-pane is fixed to the top of the stem, and the specific gravity computed from the weight necessary to be placed on this pan to cause the hydrometer to sink to the same point in all liquids for which it is used. 461. Depth of Flotation.—If the weight and external shape of the floating body are known, and the centre of gravity situated that the position of flotation is known, the depth of the lowest point below the surface may be determined. **FLotation.** 503 **Case I. Right prism or cylinder with its axis vertical.** Fig. 505. (For stability in this position, see § 464a.) Let $G'$ = weight of cylinder, $F$ the area of its cross-section (full circle), $h'$ its altitude, and $h$ the unknown depth of flotation (or draught); then from eq. (1), § 426, $$G' = Fh'; \therefore h = \frac{G'}{F}; \quad (1)$$ in which $y$ = heaviness of the liquid. If the prism (or cylinder) is homogeneous (and then $C$, at the middle of $h'$, is higher than $B$) and $\gamma'$ its heaviness, we then have $$h = \frac{F\gamma'h'}{F\gamma} = \frac{\gamma'h'}{\gamma} = \sigma h'; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)$$ in which $\sigma$ = specific gravity of solid referred to the liquid as standard. (See § 458.) **Case II. Pyramid or cone with axis vertical and vertex down.** Fig. 506. Let $V' =$ volume of whole pyramid (or cone), and $V' =$ volume of displacement. From similar pyramids, $$\frac{V}{V'} = \frac{h'}{h}; \therefore h = \sqrt[3]{\frac{V}{V'}}; h'.$$ But $G' = V'y$; or, $V = \frac{G}{y}$; whence $$h = h'\sqrt[3]{\frac{G}{V'y}} \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad (3)$$ 594 MECHANICS OF ENGINEERING. CASE III. Ditto, but vertex up.—Fig. 507. Let the notation be as before, for $V$ and $V'$. The part out of water is a pyramid of volume $V'' = V' - V$, and is similar to the whole pyramid; Also, $$V = \frac{G'}{r}$$ ∴ $h'' = h'\sqrt{\frac{V''}{V'}} = h'\sqrt{\frac{V'' - G'}{V'}}$$ ∴, finally, $h = h\left[1 - \sqrt{1 - (G' + V')}\right]$ (6) CASE IV. Sphere—Fig. 508. The volume immersed is $$V = \int_{a}^{b} \pi r^2 ds = \pi \int_{a}^{b} (2ra - x^2) dx = \pi h\left[\frac{A}{3}\right]$$ and hence, since $Vr = G' = \text{weight of sphere}$, $$\pi rh - \frac{\pi h^3}{3} = \frac{G'}{r}$$ (5) From which cubic equation $h$ may be obtained by successive trials and approximations. (A) The first solution of (5) for the unknown $h$ is impossible, as it fails under the irreducible case of Cardan's Rule.) CASE V. Right cylinder with axis horizontal.—Fig. 509. Vol. im- mersa. $= V' = (\text{area of seg. } ADE)\times l$ $$= (r'a - 4r^2\sin 2a);$$ ∴ hence, since $V = \frac{G'}{r}$ $$l^2[a - 4\sin 2a] = \frac{G'}{r}$$ (6) Fig. 507 Fig. 508 Fig. 509 EXAMPLES OF FLOTATION. From this transcendental equation we can obtain $\alpha$, by trial, in radians (see example in § 428), and finally $A$, since $$A = r(1 - \cos x). \qquad (7)$$ **Example 1.**—A sphere of 40 inches diameter is observed to have a depth of flotation $A = 9$ in. in water. Required its weight $G'$. From eq. (6) (inch, lb., sec.) we have $$G' = [62.5 + 1728] \times 9[30 - \frac{1}{4} \times 9] = 156.5 \text{ lbs}.$$ The sphere may be hollow, e.g., of sheet metal loaded with shot; constructed in any way, so long as $G'$ and the volume $V$ of displacement remain unchanged. But if the sphere is homogeneous, its heaviness $(\text{or } G')$ must be $$= G' + V = G' + \pi r^2 A = (156.5 + \frac{1}{4}\pi 20^2)$$ = .00466 lbs. per cubic inch, and hence, referred to water, its specific gravity is $\sigma =$ about .018. **Example 2.**—The right cylinder in Fig. 509 is homogeneous and 10 inches in diameter, and has a specific gravity (referred to water) of $\sigma = 0.30$. Required the depth of flotation $A$. Its heaviness must be $y' = \sigma y$; hence its weight $$G' = V \sigma y = \pi r l y \sigma;$$ hence, from eq. (6), $$r^2[l(\alpha - \frac{1}{4}\sin 2a) = \pi r l \sigma, i.e. -\frac{1}{4}\sin 2a = \pi \sigma$$ (involving abstract numbers only). Trying $\alpha = 60^\circ (= \frac{1}{4}\pi$ in radians), we have $$\frac{1}{4}\pi - \frac{1}{4}\sin 190^\circ = 0.614; whereas \pi \sigma = .9424$$ For $\alpha = 70^\circ$, $1.9217 - \frac{1}{4}\sin 140^\circ = .9008$; For $\alpha = 71^\circ$, $1.3391 - \frac{1}{4}\sin 142^\circ = .9831$; For $\alpha = 72^\circ$, $1.3455 - \frac{1}{4}\sin 143^\circ = .9428$, which may be considered sufficiently close. Now from eq. (7), $$A = (5 \text{ in.}) (1 - \cos 71^\circ) (22') = 3.40 \text{ in.}$$ 506 MECHANICS OF ENGINEERING. 482. Draught of Ships.—In designing a ship, especially if of a new model, the position of the centre of gravity is found by eq. (3) of § 25 (with weights instead of volumes); i.e., the sum of the products obtained by multiplying the weight of each portion of the hull and its distance from the centre of gravity from a convenient reference-plane (e.g., the horizontal plane of the keel bottom) is divided by the sum of the weights, and the quotient is the distance of the centre of gravity of the whole from the reference-plane. Similarly, the distance from another reference-plane is determined. These two co-ordinates and the fact that the centre of gravity lies in the median vertical plane of symmetry of the ship (assuming a symmetrical arrangement of the framework and cargo) give its location. The total volume, of course, is the sum of the individual weights just mentioned. The centre of buoyancy, for any assumed draught and corresponding position of ship, is found by the same method; but more simply, since it is the centre of gravity of the imaginary homogeneous volume between the water-line plane and the vertical surface of the hull. This volume (of "displacement") is divided into an even number (say 4 to 8) of horizontal slices, and Simpson's Rule applied to find the volume $V$ of these slices for any assumed draught, § 28; $h$ is the height of its centre of gravity above the keel. Similarly, by division into (from 8 to 20) vertical slices, $7$ to keel (an even number and of equal thickness), we find the distance of the centre of gravity from the bow. Thus the centre of buoyancy is fixed, and the corresponding buoyant effort $F_y$ (technically called the displacement and usually expressed in tons) computed, for any assumed draught of ship (upright). That position at which this effort acts vertically below the water line is the position of equilibrium of the ship when floating upright in still water, and the corresponding draught is noted. As to whether this equilibrium is stable or unstable, the following will show. In most ships the centre of gravity $C$ is several feet above the centre of buoyancy $B$, and a foot or more below the water line. DRAUGHT OF SHIPS. 507 After a ship is afloat and its draught actually noted its total weight $G' = V_y$, can be computed, the values of $V$ for different draughts having been calculated in advance. In this way the weights of different cargoes can also be measured. EXAMPLE.—A ship having a displacement of 5000 tons is itself 5000 tons in weight, and displaces a volume of salt water $V = G' + y = 10,000,000 \text{ lbs.} - 64 \text{ lbs. per cub. ft.} = 156250$ cub. ft. 463. Angular Stability of Ships.—If a vessel floating upright were of the peculiar form and position of Fig. 510 (the water-line section having an area = zero) its tendency to regain that position, or depart from it, when slightly inclined an angle $\phi$ from the vertical is due to the action of the couple now formed by the equal and parallel forces $V_y$ and $G'$ which act at right angles to each other. This couple is called a righting couple if it acts to restore the first position (as in Fig. 511, where $C$ is lower than $B$), and an upsetting couple if the reverse, $C$ above $B$. In either case the moment of the couple is $$= V_y \cdot BC \sin \phi = V_y c \sin \phi,$$ and the centre of buoyancy $B$ does not change its position in the vessel, since the water-displacing shape remains the same; i.e., no new portions of the vessel are either immersed or raised out of the water. But in a vessel of ordinary form, when turned an angle $\phi$ from the vertical, Fig. 512 (in which $ED$ is a line which is vertical when the ship is upright), there is a new centre of buoyancy $B$, corresponding to the new shape $A', N', D'$ of the displacement-volume, and the couple to right the vessel (or the reverse) A diagram showing a ship with a waterline section having an area of zero, and two forces acting on it: one force is perpendicular to the waterline section and acts at point C, while the other force acts at point B. The diagram shows how these forces create a righting couple. Fig. 510. A diagram showing a ship with a waterline section having an area of zero, and two forces acting on it: one force is perpendicular to the waterline section and acts at point C, while the other force acts at point B. The diagram shows how these forces create an upsetting couple. Fig. 511. A diagram showing a ship with a waterline section having an area of zero, and two forces acting on it: one force is perpendicular to the waterline section and acts at point C, while the other force acts at point B. The diagram shows how these forces create a righting couple. Fig. 512. 508 MECHANICS OF ENGINEERING. consists of the two forces $G$ at $C$ and $V_Y$ at $B_1$, and has a moment (which we may call $M$, or moment of stability) of a value (§ 28) $$M = V_Y \cdot m C \sin \phi \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . \quad (1)$$ Now conceive put in at $B$ (centre of buoyancy of the upright position) two vertical and opposite forces, each = $V_Y = G'$, calling them $P$ and $P'$ (see § 90), Fig. 512. We can now regard the couple ($G', V_Y$) as replaced by the two couples ($G', P$) and ($P', V_Y$); for evidently $$V_Y \cdot m C \sin \phi = V_Y \cdot BC \sin \phi + V_Y \cdot m B \sin \phi;$$ (§§ 83 and 84.) $$\therefore M = V_Y \cdot BC \sin \phi + V_Y \cdot m B \sin \phi \quad. .. (2)$$ But the couple ($G', P$) would be the only one to right the vessel if no new portions of the hull entered the water or emerged from it, in the inclined position; hence the other couple ($P', V_Y$) owes its existence to the emission of the wedge $A O_A$, and the immersion of the wedge $NO_N$; i.e., to the loss of volume (vol-ume $A O_A$) $\times Y$ on one side, and the gain of an equal buoyant force on the other; therefore this couple ($P', V_Y$) is the equivalent of the couple ($Q', Q$), Fig. 51C, formed by putting in at the centre of buoyancy of each of the two wedges a vertical force $$Q = (\text{vol. of wedge}) \times Y = V_{w.Y}.$$ (See figure.) Fig. 513. Fig. 514. STABILITY OF SHIPS. 590 If $a$ denotes the arm of this couple, we may write $$V_Y \cdot mB \sin \phi, [of eq. (9)], = V_{\text{ws}}; \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ and hence, denoting $BC$ by $b$, we have $$M = \pm V_{\text{ws}} \sin \phi + V_{\text{ws}}; \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. (4)$$ the negative sign in which is to be used when $C$ is above $B$ (as with most ships). $O$, the intersection of $ED$ and $AN$, does not necessarily lie on the new water-line plane $X,N$. EXAMPLE.--If a ship of $(FV) = 3000$ tons displacement with $C$ 4 ft. above $B$ (i.e., $\phi = 0$) is deviated 10° from the vertical, its centre of buoyancy, for which angles of inclination are considered, will be at $NON$, having each a volume of 4000 cubic feet, while the horizontal distance between their centres of buoyancy is 18 feet, the moment of the setting couple will be, from eq. (4) (ft-ton.-sec system, in which $y$ of salt water = 0.032), $$M = -3000\times4\times0.1736 + 4000\times0.032\times18 = 220.8\,\text{ft.\ ton},$$ which being + indicates a righting couple. 484. Remark.--If with a given ship and cargo this moment of stability, $M$, be computed, by eq. (4), for a number of values of $\phi$, and the results plotted as ordinates (to scale) of a curve, $\phi$ being the abscissa, the curve obtained is indicative of the general stability of the ship. See Fig. 514. For some value of $\phi = OK$ (as well as for $\phi = 0$) the value of $M$ is zero, and for $\phi > OK$, $M$ is negative, indicating an upsetting couple. That is, for $\phi = 0$ the equilibrium is stable, but for $\phi = OK$, unstable ; and $M = 0$ in both positions. From eq. (4) we see why, if $C$ is above $B$, instability does not necessarily follow. 484a. Metacentre of a Ship.--Returning again to Fig. 512, we note that the entire couple $(G,V_Y)$ will be a righting couple, or an upsetting couple, according as the point $m$ (the A diagram showing the metacentre \( M \) and the center of gravity \( G \) of a ship, with \( V_Y \) representing the vertical distance from \( G \) to the waterline.) Fig. 514 600 MECHANICS OF ENGINEERING. intersection of the vertical through $B$, the new centre of buoyancy with $BC$ prolonged) is above or below the centre of gravity $C$ of the ship. The location of this point $m$ changes with $\phi$; but as $\phi$ becomes very small (and ultimately zero) $m$ approaches a definite position on the line $DE$, though not occurring at that point. This point $m$ is called the metacentre, and accordingly the following may be stated: A ship floating upright is in stable equilibrium if its metacentre is above its centre of gravity; and vice versa. In other words, for a slight inclination from the vertical a righting, and not an upsetting, couple is called into action, if $m$ is above $C$. To find the metacentre, by means of the distance $BM$, we have, from eq. (3): $$\frac{m}{B} = \frac{V_{y} \cdot yA}{V_y \sin \phi}, \quad \ldots \ldots \ldots (5)$$ and wish ultimately to make $\phi = 0$. Now the moment $(V_y)a = z$ the sum of the moments about the horizontal fore-and-aft water-line axis $OL$, Fig. 515, of the buoyant efforts due to the immersion of the separate vertical elementary prisms of the wedge $OLY_1N_1$, plus the moments of those lost from immersion, in the wedge $O_2L_2N_2$. Let $a_1a_2a_3a_4a_5$ be the new water-line section of the ship when inclined a small angle $\phi$ from the vertical ($\phi = NO,N_1$, and $OALR$ the old water-line). Let $z$ be the distance of any elementary area $dF$ of the water-line section from $OL$ (which is the intersection of the two water-line planes). Each $dF$ is the base of an elementary prism with altitude $y$, so that the wedge $NO,LN_1$ (or of wedge $AOLR$ when $\phi$ is negative). The volume element is $(\text{vol}) \times y = ydF$, and its moment about $OL$ is $\phi ydF$. Hence the total moment, $= Qa$, or $V_ya$, by Fig. 518. $$= \phi y^2 dF = y\phi \times I_{ox}$$ THE METACENTRE. 601 of water-line section, in which $I_{\text{ok}}$ denotes the "moment of inertia" (§ 85) of the plane figure $OLNO$ about the axis $OL$. Hence from (5), putting $\phi = \sin \phi$ (true when $\phi = 0$), we have $mC = I_{\text{ok}} + V$, and therefore the distance $mC$ of the meta-centre $m$ above $C$, the centre of gravity of the ship, Fig. 512, is $$\frac{mC}{V} = h_m = \frac{I_{\text{ok}} (\text{of water-line sec})}{V} \pm e, \dots \dots (6)$$ in which $e = BC =$ distance from the centre of gravity to the centre of buoyancy, the negative sign being used when $C$ is above $B$; while $V =$ whole volume of water displaced by the ship. We may also write, from eqs. (6) and (1), for small values of $\phi$, $$\text{Mom. of righting couple} = M = VY \sin \phi \left[ \frac{I_{\text{ok}}}{V} \pm e \right], \dots (7)$$ or $$M = y \sin \phi (I_{\text{ok}} + Ve). \dots (7')$$ Eqs. (7) and (7') will give close approximations for $\phi < 10^\circ$ or $15^\circ$ with ships of ordinary forms. EXAMPLE 1. A homogeneous right parallelopiped, of heaviness $y'$, floats upright in air, Fig. 513. Find the distance $e$ of $m$ from $C$, for its metacentre in this position, and whether the equilibrium is stable. Here the centre of gravity, $C$, being the centre of figure, is, of course above $B$, the centre of buoyancy; hence $e$ is negative. $B$ is the centre of gravity of the displacement, and is therefore a distance $\frac{l}{2}$ below the water-line. We here assume that $l$ is greater than $b'$. From eq. (3), § 461, $$A = \frac{k'Y'}{y'};$$ 602 MECHANICS OF ENGINEERING. and since $CD = \frac{1}{2}h'$, and $BD = \frac{1}{2}h$, i.e. $e = \frac{1}{2}(N - h)$; i.e., $e = \frac{1}{2}\left[1 - \frac{r'}{r}\right]$; while (§ 90) $I_{ok}$, of the water-line section $AN_1 = \frac{1}{2}bh''$. Also, $$V = b\frac{h'}{r} = b\frac{h''}{r'}$$ and hence, from eq. (6), we have $$A_m = \frac{b'h''}{12h''r'} - \frac{1}{4}\left[1 - \frac{r'}{r}\right] = \frac{b''-0h''(1-\frac{r'}{r})}{12h''r'}$$ Hence if $b'' > 6h''\frac{r'}{r}(1-\frac{r'}{r})$, the position in Fig. 516 is one of stable equilibrium, and vice versa. E.g., if $y' = 4y$, $v = 12$ inches and $h' = 6$ inches, we have (inch, pound, sec.) $$A_m = mC = \frac{1}{4}\left[\frac{144-6\times4(1-4)}{3}\right] = 2.5\;\mathrm{in}.$$ The equilibrium will be unstable if, with $y' = 4y$, $b''$ is made less than 1.925 $h'$; for putting $mC = 0$, we obtain $b' = 1.225 A'$. EXAMPLE 2.—(Ft., lb., sec.) Let Fig. 517 represent the half water-line section of a loaded ship of $G' = V'y' = 1010$ tons displacement; required the height of the metacentre above the centre of buoyancy, i.e., $mC = 0$. (See equation just before eq. (6).) Now the quantity $I_{ok}$, of the water-line section, may, from symmetry, (see § 93,) be written $$I_{ok} = 2\int_0^{\pi/2}k'y'dx\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;(\text{(i)})$$ METACENTRE. 603 in which $y$ = the ordinate $\gamma$ to the axis $OL$ at any point; and this, again, by Simpson's Rule for approximate integration, $OL$ being divided into an even number, $n$, of equal parts, and ordinates erected (see figure), may be written $$I_{ac} = \frac{9}{8} \cdot \frac{\partial L - 0}{\partial n} \left[ y_1^2 + 4(y_1^2 + y_2^2 + \cdots + y_n^2) \right.$$ $$+ 2(y_1^2 + y_2^2 + \cdots + y_{n-1}^2) + y_n^2 \right].$$ From which, by numerical substitution (see figure for dimension; $n=8$), $$I_{ac} = \frac{9}{8} \cdot 160 \left[ (0.5^2 + 4(0.5^2 + 1.5^2 + 1.7^2)) \right.$$ $$+ 2(0.9^2 + 1.4^2 + 1.1^2) + 0.5^2 \right];$$ or, $$I_{ac} = 1738 \cdot 729$$ $$= 19197 \cdot 8184$$ $$= 943 \cdot 1331$$ $$= 120801 \text{ biquad. ft.; } .mB = I_{ac} = 120801$$ $$= 3.8 \text{ feet}.$$ That is, the metacentre is 3.8 feet above the centre of buoyancy, and hence, if $BC = 3$ feet, is 1.90 ft. above the centre of gravity. [See Johnson's Cyclopaedia, article Naval Architecture.] 465. Metacentre for Longitudinal Stability.—If we consider the stability of a ship when subject to pitching, in a manner similar to that just pursued for rolling, we derive the position of the metacentre for pitching or for longitudinal stability— and this of course occupies a much higher position than that for rolling, involving as it does the moment of inertia of the water-line section about a horizontal gravity axis $\gamma$ to the keel. With this one change, eq. (6) holds for this case also. In large ships the height of this metacentre above the centre of gravity of the ship may be as great as 30 feet. A diagram showing the metacentre and its relationship to the centre of gravity and buoyancy. CHAPTER V. HYDROSTATICS (Continued)---GASEOUS FLUIDS. 486. Thermometers.--The temperature, or "hotness," of liquids has, within certain limits, but little influence on their statical behavior, but with gases must always be taken into account, since the three quantities, tension, temperature, and volume, of a given mass of gas are connected by a nearly invariable law, as will be seen. An air-thermometer, Fig. 518, consists of a large glass bulb filled with air, from which projects a fine straight tube of even bore (so that equal lengths represent equal volumes). A small drop of liquid, $A$, separates the internal air from the external air, both of which are at a tension of (say) one atmosphere (14.7 lbs. per sq. inch). When the bulb is placed in melting ice (freezing-point) the drop is found at point $F$ in the tube; when in boiling water (boiling under a pressure of one atmosphere), the drop is found at $B$, on account of the expansion of the internal air under the influence of the heat imparted to it. (The glass also expands, but only about $\frac{1}{10}$ as much; this will be neglected.) The distance $FB$ along the tube may now be divided into a convenient number of equal parts called degrees. If into one hundred degrees, it is found that each degree represents a volume equal to the $\frac{1}{1000}$ (0.001) part of the volume at freezing-point; i.e., the increase of volume from the temperature of freezing-point to that of the boiling point of water = 0.367 of the volume at freezing, the pressure being the same, and even having any value whatever (as well as one atmosphere), within ordinary limits, so long as it is the same both at freezing and boiling. A diagram showing an air-thermometer with a glass bulb filled with air and a fine straight tube projecting from it. The tube is divided into equal parts representing different temperatures. 604 THERMOMETERS. 605 ing. It must be understood, however, that by temperature of boiling is always meant that of water boiling under one atmosphere pressure. Another way of stating the above, if one hundred degrees are used between freezing and boiling, is as follows: That for each degree increase of temperature the increase of volume is 273/273 = .00867 total volume at freezing; 273 being the reciprocal of .00867. As it is not always practicable to preserve the pressure constant under all circumstances with an air-thermometer, we use the common mercurial thermometer for most practical purposes. In this, the tube is sealed at the outer extremity, with a vacuum above the column of mercury, and its indications agree very closely with those of the air-thermometer. That equal absolute increments of volume should imply equal increments of heat has been shown by experiments, though it may be said (for constant pressure) could not reasonably be asserted without satisfactory experimental evidence. This, however, is not altogether wanting, so that we are enabled to say that within a moderate range of temperature equal increments of heat produce equal increments of volume in a given mass not only of atmospheric air, but of the so-called "perfect" or "permanent" gases, oxygen, nitrogen, hydrogen, etc. (so named before it was found that they could be liquefied). This is nearly true for mercury, which is not so perfect for heat. Alcohol has never been frozen, and hence is used instead of mercury as a thermometric substance to measure temperatures below the freezing-point of the latter. The scale of a mercurial thermometer is fixed ; but with an air-thermometer we should have to use a new scale, and in a new position on the tube, for each value of the pressure. 467. Thermometric Scales.—In the Fahrenheit scale the tube between freezing and boiling is marked off into 180 equal parts, and the zero placed at 32 of these parts below the freezing-point, which is hence +32°, and the boiling-point +212°. The Centigrade, or Celsius, scale, which is the one chiefly used in scientific practice, places its zero at freezing, and 100 at boiling-point. Hence to reduce 606 MECHANICS OF ENGINEERING. Fahr. readings to Centigrade, subtract 33° and multiply by 4; Cent. Fahrenheit, multiply by 4 and add 33°. 468. Absolute Temperature.—Experiment also shows that if a mass of air or other perfect gas is confined in a vessel whose volume is but slightly affected by changes of temperature, equal increments of heat imparted to the gas produce equal increments of tension (i.e., pressure per unit area); or, as to the amount of the increase, that when the temperature is raised by an amount 1° Centigrade, the tension is increased $\frac{1}{4}$ of its value at freezing-point. Hence, theoretically, an ideal barometer (containing a liquid unaffected by changes of temperature) communicating with the confined gas (whose volume practically remains constant) would be a perfect barometer, Fig. 519, and the attached scale could be graduated accordingly. Thus, if the column stood at A when the temperature was freezing, $A$ would be marked 0° on the Centigrade system, and the degree spaces above and below $A$ would each = $\frac{1}{4}$ of the height $AB$, and therefore the point $B$ (cistern level) to which the column would sink if the gas-tension were zero would be marked —273° Centigrade. But a zero-pressure, in the Kinetic Theory of Gases (§ 408), signifies that the gaseous molecules, no longer impinging against the vessel walls (so that the press. = 0), have become motionless; and this, in the Mechanical Theory of Heat, or Thermodynamics, implies that the gas is totally destitute of heat. Hence this ideal temperature of —273° Centigrade, or —46° Fahrenheit, is called the Absolute Zero Temperature; and by reeksolding the formulae expressing the various thermodynamic formulae will render them much more simple and compact. Temperature so reckoned is called absolute temperature, and will be denoted by the letter $T$. Hence the following rules for reduction : A diagram showing a barometer with a liquid column in a vessel containing air. The liquid level is at point A when the temperature is freezing point. Fig. 519 GASES AND VAPORS. 607 Aboel. temp. $T$ in Cent. degrees = Ordinary Cent. + 273°; Aboel. temp. $T$ in Fahr. degrees = Ordinary Fahr. + 460°. For example, for 80° Cent., $T = 293°$ Abs. Cent. 469. Distinction Between Gases and Vapors.—All known gases can be converted into liquid by a sufficient reduction of temperature or increase of pressure, or both; some, however, with great difficulty, such as atmospheric air, oxygen, hydrogen, nitrogen, etc., these having been but recently (1875) reduced to the liquid form. A vapor is a gas near the point of liquefaction which does not change its physical character under changes of temperature and pressure characteristic of a gas when at a temperature much above the point of liquefaction. All gases treated in this chapter (except steam) are supposed in a condition far removed from this stage. The following will illustrate the properties of vapors. See Fig. 520. Let a quantity of liquid, say water, be introduced into a closed space previously vacuum, of considerably larger volume than the water, and furnished with a manometer. Vapor of water immediately begins to form in the space above the liquid, and continues to do so until its pressure attains a definite value dependent on the temperature, and not on the ratio of the volume of the vessel and the original volume of water; e.g., if the temperature is 70° Fahrenheit, the vapor ceases to form when the tension reaches a value of 0.36 lbs. per sq. inch. If heat is gradually applied to raise the temperature, the vapor forms for a time from the body of the liquid, unless the heat is applied very slowly), but the tension will not rise above a fixed value for each temperature (independent of size of vessel) so long as there is any liquid left. Some of these corresponding values, for water, are as follows: For a Fahr. temp. = 70° 100° 150° 212° 290° 287° 300° Tension (lbs.) = 0.36 0.93 3.69 14.7 17.2 55.0 67.2 per sq.in.) At any such stage the vapor is said to be saturated. A diagram showing a closed space with a manometer attached to it, filled with water and vapor. Fig. 520 608 MECHANICS OF ENGINEERING. Finally, at some temperature, dependent on the ratio of the original volume of water to that of the vessel, all of the water will have been converted into vapor (i.e., steam); and if the temperature be still further increased, the tension also increases and no longer corresponds to the temperature alone, but also on the heaviness of the vapor when this is disregarded. The vapor is now said to be superheated, and conforms more in its properties to perfect gases. 470. Critical Temperature.—From certain experiments there seems to be reason to believe that at a certain temperature, called the critical temperature, different for different liquids, all of the liquid in the vessel (if any remains), and supposing the vessel to be large enough, will be completely converted into vapor, whatever be the size of the vessel. That is, above the critical temperature the substance is necessarily gaseous, in the most exclusive sense, incapable of liquefaction by pressure alone; while below this temperature it is a vapor, and liquefaction will begin if, by compression in a cylinder and consequent increase of pressure, the tension can be raised to a value corresponding, for a state of saturation, to the temperature (in such a table) as just given for water. For example, if vapor of water at 39° Fahrenheit occupies 1 lb. per sq. inch (this is superheated steam since 290° is higher than the temperature which for saturation corresponds to p = 10 lbs. per sq. inch) is compressed slowly (slowly, to avoid change of temperature) till the tension rises to 17.2 lbs. per sq. in., which (see above table) is the pressure of saturation for a temperature of 250° Fahrenheit for water vapor, the vapor is saturated, i.e., liquefaction is ready to begin, and during any further reduction of volume the pressure remains constant and so does the variation of tension. By "perfect gases" or gases proper, we may understand, therefore, those which cannot be liquefied by pressure unaccompanied by great reduction of temperature; i.e., whose "critical temperatures" are very low. The critical temperature of N₂O or nitrous oxide gas, is between — 11° and — 8° Centigrade, while that of oxygen is said to be at — 118° Centi- LAW OF CHARLES. 609 grade. [See p. 471, vol. 129 of the *Journal of the Franklin Institute*. For an account of the liquefaction of oxygen, etc., see the same periodical, January to June, 1878.] **471. Law of Charles (and of Gay Lussac).—The mode of graduation of the air-thermometer may be expressed in the following formula, which holds good (for practical purposes) within the ordinary limits of experiment for a given mass of any perfect gas, the tension remaining constant: $$V = V_s + 0.00367 V_s \cdot t = V_s(1 + 0.0367t); \quad \ldots \quad (1)$$ in which $V_s$ denotes the volume occupied by the given mass at freezing-point under the given pressure, $V$ its volume at any other temperature $t$ Centigrade under the same pressure. Now, 273 being the reciprocal of 0.00367, we may write $$V = V_s \frac{(273 + t)}{273}; \text{i.e., } \frac{V}{V_s} = \frac{T}{T_s} \quad \ldots \quad (\text{const.}) \quad (9)$$ (see § 485) in which $T_s$ = the absolute temperature of freezing-point, = 273° absolute Centigrade, and $T$ the absolute temperature corresponding to $t$ Centigrade. Eq. (9) is also true when $T$ is negative, both expressions taking Fahrenheit degrees (from absolute zero, of course). Accordingly, we may say that, the pressure remaining the same, the volume of a given mass of gas varies directly as the absolute temperature. Since the weight of the given mass of gas is invariable at a given place on the earth's surface, we may always use the equation $$V_Y = V_{Y_s}, \ldots \ldots \ldots \quad (8)$$ pressure constant or not, and hence (9) may be rewritten $$\frac{Y_s}{Y} = \frac{T}{T_s} \quad (\text{press. const.}) \quad (4)$$ i.e., if the pressure is constant, the heaviness (and therefore the specific gravity) varies inversely as the absolute temperature. 610 MECHANICS OF ENGINEERING. Experiment also shows (§ 468) that if the volume [and therefore the heaviness, eq. (3)] remains constant, while the temperature varies, the tension $p$ will change according to the following relation, in which $p_t$ = the tension when the temperature is freezing: $$p = p_t + \frac{273 + t}{278} p_t = p_t \left(1 + \frac{t}{278}\right), \dots \dots \quad (5)$$ $t$ denoting the Centigrade temperature. Hence transforming, as before, we have $$\frac{p}{p_t} = \frac{T}{T_t} \cdot \left\{\begin{array}{l} {\text{vol., and } c.}; \\ {\text{leav., const.}} \end{array}\right.; \quad (6)$$ or, the volume and heaviness remaining constant, the tension of a given mass of gas varies directly as the absolute temperature. This is called the Law of Charles (or of Gay Lussac). 472. General Formula for any Change of State of a Perfect Gas. If any two of the three quantities, viz., volume (or heaviness), tension, and temperature, are changed, the new value of the third is determinate from those of the other two, according to a relation proved as follows (remembering that henceforth the absolute temperature only will be used, § 468): Fig. 521. At A a certain mass of gas at a tension of $p$, one atmosphere, and absolute temperature $T$, (freezing), occupies a volume $V_0$, length $L_0$. At B, with same tension $= T'$, without change of tension (expanding behind a piston, for instance). Its volume will increase to a value $V$ which from (9) of 471 will satisfy the relation $$\frac{V}{V_0} = \frac{T'}{T}, \dots \dots \dots \quad (7)$$ (See B in figure.) Let it now be heated without change of volume to an absolute temperature $T' (C$ in figure). Its volume is still $V$, but Laws of perfect gases. 611 the tension has risen to a value $p$, such that, on comparing $B$ and $C$ by eq. (6), we have $$\frac{P}{P_0} = \frac{T}{T_0}. \qquad \qquad (8)$$ Combining (7) and (8), we obtain for any state in which the tension is $p$, volume $V$, and absolute temperature $T$, in (General) . . . . $\frac{PV}{T} = \frac{P_0V}{T_0}$; or $\frac{PV}{T} = \text{a constant};$. (9) or (General) . . . . $\frac{p_nV_n}{T_n} = \frac{p_sV_s}{T_s}, \qquad \qquad (10)$ which, since (General) . . . $V_N = V_{N_s} = V_{N_m} = V_{N_s}, \qquad (11)$ is true for any change of state, we may also write (General) . . . . $\frac{p}{yT} = \frac{p_0}{y_0T_0}, \qquad \qquad (12)$ or $\frac{p_n}{y_nT_n} = \frac{p_s}{y_sT_s}. \qquad \qquad (13)$ These equations (9) to (13), inclusive, hold good for any state of a mass of any perfect gas (most accurately for air). The subscript 0 refers to the state of one-atmosphere tension and freezing-point temperature, $m$ and $n$ to any two states whatever (within practical limits); $y$ is the heaviness, §§ 7 and 409, and $T$ the absolute temperature, § 468. If $p$, $V$, and $T$ of equation (9) be treated as variables, and laid off to scale as co-ordinates parallel to three axes in space, respectively, the surface so formed of which (9) is the equation is a hyperbolic paraboloid. 473. Examples.—Example 1. What cubic space will be occupied by 2 lbs. of hydrogen gas at a tension of two atmospheres and a temperature of $27^\circ$ Centigrade? A diagram showing the relationship between pressure, volume, and temperature for a perfect gas. 612 MECHANICS OF ENGINEERING. With the inch-lb.-sec. system we have $p_1 = 14.7$ lbs. per sq. inch, $y_1 = [0.058 + 1728]$ lbs. per cubic inch, and $T_1 = 273^\circ$ absolute Centigrade, when the gas is at freezing-point at one atmosphere (i.e., in state sub-zero). In the state mentioned in the problem, we have $p = 2 \times 14.7$ lbs. per sq. in., $$T = 273 + T_1 = 300^\circ$$ absolute Centigrade, while $y$ is required. Hence, from eq. (12), $$\frac{2 \times 14.7}{y \times 300} = \frac{14.7}{(0.058 + 1728) \times 273};$$ :. $y = \frac{0.012}{1728}$ lbs. per cub. in. = $\frac{0.012}{1}$ lbs. per cub. foot; and if the total weight, $G$, is to be 2 lbs., we have $(f_1 \text{ lb. sec.) } V = \frac{2}{0.012} = 196$ cubic feet.--Ans. Example 2.--The mass of air originally at 94° Centigrade and a tension indicated by a barometric column of 40 inches of mercury has been similarly reduced to half its former volume and heated to 100° Centigrade; required its tension in this new state, which we call the state $n$, $m$ being the original state. Use the inch, lb., sec. We have given, therefore, $$p_n = \frac{48}{4} \times 14.7 \text{ per sq. inch}, T_n = 273 + 24 = 297^\circ$$ absolute Centigrade, the ratio $$V_m : V_n = 2 : 1,$$ and $T_n = 273 + 100 = 373^\circ$ Abs. Cent.; while $p_n$ is the unknown quantity. From eq. (10), hence, $$p_n = \frac{V_m}{V_n} \cdot \frac{T_n}{T_m} \cdot p_m = 2 \times \frac{48}{4} \times 14.7 = 49.92 \text{ lbs. per sq. in.,}$$ which an ordinary steam-gauge would indicate as $$(49.92 - 14.7) = 34.52 \text{ lbs. per sq. inch}. ($$ (That is, if the weather barometer indicated exactly 14.7 lbs per sq. inch.) EXAMPLES. PERFECT GASES. 613 **EXAMPLE 3.** — A mass of air, Fig. 592, occupies a rigid closed vessel at a temperature of 15° Centigrade (equal to that of surrounding objects) and a tension of four atmospheres [state $m$]. By opening a stop-cock for a few seconds, thus allowing a portion of the gas to escape quickly, and then shutting it, the remainder of the air (now in state $n$) is found to have a tension of only 2.5 atmospheres (measured immediately); its temperature cannot be measured immediately (so much time being necessary to affect a thermometer), and is less than before. To compute this temperature, $T_n$, we allow the air now in the vessel to come again to the same temperature as surrounding objects (15° Centigrade); find then the tension to be 2.92 atmospheres. Call the last state, state $r$ (incl. lb., see.). The problem then stands thus: $$p_m = 4 \times 14.7$$ $$T_m = 288^\circ\mathrm{Abe.\ Cent.}$$ $$p_n = 2.5 \times 14.7$$ $$T_n = ?$$ In states $n$ and $r$ the heaviness is the same; hence an equation like (6) of § 471 is applicable, whence $$\frac{P_r}{P_n} = \frac{T_r}{T_n}, \quad \text{or} \quad T_n = \frac{2.5 \times 14.7}{2.92} \times 288 = 346^\circ\mathrm{Abe.\ Cent.}$$ or - 27° Centigrade; considerably below freezing, as a result of allowing the sudden escape of a portion of the air, and the consequent sudden expansion, and reduction of tension, of the remainder. In this sudden passage from state $m$ to state $n$, the remainder altered its heaviness (and its volume in inverse ratio) in the ratio (see eqs. (11) and (10) of § 472) $$\frac{V_r}{V_n} = \frac{P_r}{P_n} \cdot \frac{T_r}{T_n} = \frac{2.5 \times 14.7}{4 \times 14.7} \cdot \frac{288}{346} = 0.73.$$ Now the heaviness in state $m$ (see eq. (12), § 472) was Fig. 592. 614 MECHANICS OF ENGINEERING. $$y_{a} = \frac{p_a}{F_a}, \quad y_a T_a = \frac{4 \times 14.7 - 0.087 \times 273}{288} = \frac{306}{1728} = 0.015$$ lbs. per cub. in. = 300 lbs. per cub. ft. $$\therefore y_a = 0.73 \times y_a = 0.323 \text{ lbs. per cub. ft.},$$ and also, since $V_a = 0.73 V_a$, about $\frac{A_{\text{new}}}{A_{\text{old}}}$ of the original quantity of air in vessel has escaped. [Note.—By numerous experiments like this, the law of cooling, when a mass of gas is allowed to expand suddenly (as, e.g., behind a piston, doing work) has been determined; and vice versa, the law of heating under sudden compression; see § 487.] 474. The Closed Air-Manometer.—If a manometer be formed of a straight tube of glass, of uniform cylindrical bore, which is partially filled with mercury and then inverted in a cistern of mercury, a quantity of air having been left between the mercury and the upper end of the tube, which is closed, the tension of this quantity of air (to be computed from its observed volume and temperature) must be added to that due to the mercury column, in order to obtain the tension $p'$ to be measured. See Fig. 523. The advantage of this kind of instrument is, that to measure great tensions the tube need not be very long. Let the temperature $T$, of whole instrument, and the tension $p'$ of the air gas in the cistern, be known when the mercury in the tube stands at the same level as that in the cistern. The tension of the air in the tube must now be $p$, also, its temperature $T$, and its volume is $V' = F'h$, $F$ being the sectional area of the bore of the tube; see on left of figure. When the instrument is used, gas of unknown tension $p'$ is admitted to the cistern, the temperature of the whole instrument being noted ($= T$), and the heights $h$ and $h'$ are observed ($h + h'$ cannot be put = $h$, A diagram showing a closed air-manometer with a tube inverted into a cistern of mercury. Fig. 523 CLOSED AIR-MANOMETER. 615 unless the cistern is very large. $p'$ is then computed as follows (eq. (3), § 413): $$p' = h''y_a + p; \ldots \ldots \ldots (1)$$ in which $p$ = the tension of the air in the tube, and $y_a$ = the bisection of mercury. But from eq. (10), § 472, putting $V_1 = Fh$, and $V = F'h$, $$p' = p\frac{V}{V'}\cdot\frac{T}{T'} = \frac{h}{A}\frac{T}{T'}, p'. \ldots \ldots (2)$$ Hence finally, from (1) and (9), $$p' = h''y_a + \frac{h}{A}\cdot\frac{T}{T'}p'. \ldots \ldots (3)$$ Since $T$, $p$, and $h$, are fixed constants for each instrument, we may, from (3), compute $p'$ for any observed values of $h$ and $T$ (N.B. $T$ and $T'$ are absolute temperatures), and construct a series of tables each of which shall give values of $p'$ for a range of values of $h$. EXAMPLE.--Supposing the fixed constants of a closed air-manometer to be (in inch-lb.-sec. system) $p_0 = 14.7$ (or one atmosphere), $T_0 = 885^\circ$ Abs. Cent. (i.e., 12° Centigrade), and $A_0 = 4^{\prime}4^{\prime\prime} = 40$ inches; required the tension in the cistern indicated by $h' = 25$ inches and $A = 15$ inches, when the temperature is $-3^\circ$ Centigrade, or $T = 270^\circ$ Abs. Cent. For mercury, $y_a = [848.7 - 1728] / 885$ (§ 409) (though strictly it should be specially computed for the temperature, since it varies about 3000/2 of itself for each Centigrade degree). Hence, eq. (3), $$p' = \frac{25 \times 848.7 + 40 \times 270}{1728 + 15 \times 885} \times 14.7 = 12.26 + 37.13 = 49.39$$ lb. per sq. inch, or nearly $\frac{3}{4}$ atmospheres (steam-gauge would read 34.7 lbs per sq. in.). 475. Mariotte's Law, (or Boyle's') Temperature Constant; i.e., Isothermal Change.--If a mass of gas be compressed, or al... 616 MECHANICS OF ENGINEERING. lowed to expand, isothermally, i.e., without change of temperature (practically this cannot be done unless the walls of the vessel are conductors of heat, and then the motion must be slow), eq. (10) of § 472 now becomes (since $T_{a} = T_{b}$) $$\{\text{Mariotte's Law}\} \cdot V_a p_a = V_b p_b, \text{ or } \frac{p_a}{p_b} = \frac{V_a}{V_b}; \quad (1)$$ i.e., the temperature remaining unchanged, the tensions are inversely proportional to the volumes of a given mass of a perfect gas; or, the product of volume by tension is a constant quantity. Again, since $V_a y_a = V_b y_b$ for any change of state, $$\{\text{Mariotte's Law}\} \cdot \frac{p_a}{p_b} = \frac{y_a}{y_b}, \text{ or } \frac{p_a}{p_b} = \frac{y_a}{y_b}. \quad (2)$$ i.e., the pressures (or tensions are directly proportional to the first power of the heaviness, if the temperature is the same. This law, which is very closely followed by all the perfect gases, was discovered by Boyle in England and Mariotte in France more than two hundred years ago, but of course is only a particular case of the general formula, for any change of state in § 472 may be verified experimentally with several gases. For example, fig. 584, the tube $O M$ being closed at the top, while $P M$ is open, let mercury be poured in at $P$ until it reaches the level $A'B'$. The air in $OA'$ is now at a tension of one atmosphere. Let more mercury be slowly poured in at $P$, until the air confined in $O$ has been compressed to a volume $OA'' = \frac{1}{2}$ of $OA'$, and the height $B''E''$ then measured; it will be found to be 80 inches, i.e., $\frac{1}{2}$ atmospheres of the air in $OA'$ is equal to three atmospheres (corresponding to 60 inches of mercury). Again, compress the air in $O$ to its original volume (when at one atmosphere), i.e., to volume $OA'' = \frac{1}{3} OA'$, and the mercury height $B''E''$ will be 60 inches, show- ing a tension of three atmospheres in the confined air at $O$ Figure 584 584 **MARIOTTE'S LAW.** 617 inches of mercury in a barometer). It is understood that the temperature is the same, i.e., that time is given the compressed air to acquire the temperature of surrounding objects after being heated by the compression, if sudden. (Norm.—The law of decrease of steam-pressure in a steam-engine cylinder, after the piston has passed the point of "cut-off" and before it begins to expand, does not materially differ from Mariotte's law, which is often applied to the case of expanding steam; see § 470.) While Mariotte's law may be considered exact for practical purposes, it is only approximately true, the amount of the deviations being different at different temperatures. Thus, for decreasing temperatures the product $Vp$ of volume by tension becomes smaller, with most gases. EXAMPLE 1.—If a mass of compressed air expands in a cylinder to a pressure whose tension is 60 lbs. per sq. inch (45.8 by steam-passage) at the beginning of the expansion, which is supposed slow (that the temperature may not fall); then when it has doubled in volume its tension will be only 30 lbs. per sq. inch; when it has tripled in volume its tension will be only 20 lbs. per sq. inch, and so on. EXAMPLE 2. Diving-bell—Fig. 525. If the cylindrical diving-bell $AB$ is 10 ft. in height, in what depth, $h = l$, of salt water, can it be let down to the bottom? Let us suppose all the water to rise in the bell more than a distance $a$ ft. Call the horizontal sectional area $F$. The mass of air in the bell is constant, at a constant temperature. First, algebraically; at the surface this mass of air occupied a volume $V_m = Fh^n$ at a tension $p_m = 14.7 \times 144$ lbs. per sq. ft.; while at the depth mentioned it is compressed to a volume $V_n = F(h - a)^n$, and is at a tension $p_n = p_m + (h - a)y_a$, in which $y_a$ is buoyancy of sea water. Hence, from eq. (1), $$p_m Fh^n = [p_m + (h - a)y_a]F(h - a) \cdot \ldots \quad (5)$$ 618 MECHANICS OF ENGINEERING. $\therefore A = a \left[ 1 + \frac{F_m}{(h-a) r_m} \right];\quad\ldots\ldots\quad(4)$ hence, numerically, (ft., lb., sec.) $A = 4 \times \left[ 1 + \frac{14.7 \times 144}{(10-4) \times 64} \right] = 26.05$ feet. **476. Mixture of Gases.—It is sometimes stated that if a vessel is occupied by a mixture of gases (between which there is no chemical action), the tension of the mixture is equal to the sum of the pressures of each of the component gases present; or, more definitely, is equal to the sum of the pressures which the separate masses would exert if each in turn occupied it alone at the same temperature. This is a direct consequence of Mariotte's law, and may be demonstrated as follows : Let the actual tension be $p$, and the capacity of the vessel $V$. Also let $V_1$, $V_2$, etc., be the volumes actually occupied by the separate masses of gas, so that $V_1 + V_2 + \ldots = V;\quad\ldots\ldots\quad(1)$ and $p_1$, $p_2$, etc., the pressures they would individually exert when occupying the volume $V$ alone at the same temperature. Then, by Mariotte's law, $p_1 = V_1 p;\quad p_2 = V_2 p;\quad\text{etc.};\quad\ldots\quad(9)$ whence, by addition, we have $V(p_1 + p_2 + \ldots) = (V_1 + V_2 + \ldots) p;\quad\ldots\ldots\quad(3)$ i.e., $p = p_1 + p_2 + \ldots$ Of course, the same statement applies to any number of separate parts into which may imagine a mass of homogeneous gas to be divided. For numerical examples and practical questions in the solution of which this principle is useful, see p. 299, etc., Rankine's Steam-engine. (Rankine uses 0.365, where 0.367 has been used here.) BAROMETRIC LEVELLING. 619 477. Barometric Levelling.—By measuring with a barometer the tension of the atmosphere at two different levels, simul- taneously, and on a still day, the two localities not being widely separated horizontally, we may compute their vertical distance apart if the temperature of the stratum of air between them is known, being the same, or nearly so, at both stations. Since the horizontal air is different at different elevations of the vertical column between the two elevations $N$ and $M$, Fig. 596, we cannot immediately regard the whole of such a column as a free body (as was done with a liquid, § 12), but must consider a horizontal thin lamina, $L$, of thickness $= dz$ and at a distance $z$ (variable) below $M$, the level of the upper station, $N$ being the lower level at a distance $x$ from $M$. The tension acts downwards from $M$ downwards, since the lower lamina have to support a greater weight than the upper; and the heaviness $\gamma$ must also increase, proportionally to $p$, since we assume that all parts of the column are at the same temperature, thus being able to apply Mariotte's law. Let the tension and heaviness of the air at the upper base of the elementary lamina, $L$, be $p$ and $\gamma$ respectively. At the lower base, a distance $dz$ below the upper, the tension is $p + dp$. Let the area of the base of lamina be $F$; then its weight is found from $F(p + dp)$; its weight $\gamma F dz$, downward; its weight $\gamma F dz$, upward. For its equilibrium $(\text{vert. compn.}) = 0$. $\therefore F(p + dp) - Fp - \gamma F dz = 0;$ i.e., $dp = \gamma dz,$ (1) which contains three variables. But from Mariotte's law, § 475, eq. (9), if $p_*$ and $\gamma_*$ refer to the air at $X$, we may substitute $\gamma = \frac{\gamma_*}{p_*}$ and obtain, after dividing by $p_*$ to separate the variables $p$ and $\gamma_*$, Fig. 596. 620 MECHANICS OF ENGINEERING. $$\frac{p_{n}}{y_{s}} \cdot \frac{dp}{dz} = dz \quad \ldots \quad \ldots \quad (9)$$ Summing equations like (9), one for each lamina between $M$ (where $p = p_m$ and $z = 0$) and $N$ (where $p = p_n$ and $z = h$), we have $$\frac{p_n}{y_s} \cdot \int_0^h \frac{dp}{P} = \int_0^h dz;$$ i.e., $h = \frac{p_n}{y_s} \log_e \left[ \frac{p_n}{p_m} \right]$, $\ldots \quad \ldots \quad (3)$ which gives $h$, the difference of level or altitude, between $M$ and $N$, in terms of the observed tensions $p_n$ and $p_m$, and of $y_s$, the heaviness of the air at $N$, which may be computed from eq. (12), § 473, substituting from which we have finally $$h = \frac{P_1}{y_s} \cdot \frac{T_1}{T_2} \cdot \log_e \left[ \frac{p_n}{p_m} \right], \quad \ldots \quad (4)$$ in which the subscript 0 refers to freezing-point and one atmosphere tension; $T_1$ and $T_2$ are absolute temperatures. For the ratio $p_n : p_m$ we may put the equal ratio $h_a : h_m$ of the actual barometric heights which measure the tensions. The log. e, (or Napierian, or natural, or hyperbolic, log.) = (common log. base 10) × 3.89258. From § 804, $y_s$ of air = 0.08074 lbs. per cubic ft., and $p_m = 14.701$ lbs. per sq. inch.; $T_1 = 273$ A.H. Cent. If the temperature of the two stations (both in the shade) are not equal, a mean temp. = $(T_1 + T_2)/2$ may be used for $T_a$ in eq. (4), for approximate results. Eq. (4) may then be written $$h (\mathrm{in}\,\mathrm{feet}) = 96213 \frac{T_1}{T_2} \cdot \log_e \left[ \frac{p_n}{p_m} \right], \quad \ldots \quad (5)$$ The quantity $\frac{P_1}{y_s} = 96213$ ft., just substituted, is called the height of the homogeneous atmosphere, i.e., the ideal height which the atmosphere would have, if incompressible and non- BAROMETRIC LEVELLING—ADIABATIC LAW. 621 expansive like a liquid, in order to exert a pressure of 14.701 lb. per sq. inch upon its base, being throughout of a constant heaviness = 0.8076 lb. per cub. foot. By inversion of eq. (4) we may also write $$p_a e^{\frac{y_3 - y_1}{T_3 - T_1}} = p_a, \quad \ldots \ldots \quad (5)$$ where $e = 2.71828$ = the Napierian Base, which is to be raised to the power whose index is the abstract number $\frac{T_3 - T_1}{P_3 - P_1}$, and the result multiplied by $p_a$ to obtain $p_a$. **EXAMPLE.—Having observed as follows (simultaneously):** At lower station $N$, $A_n = 80.05$ in. mercury; temp. = 77.6° F.; “ upper “ $M$, $A_m = 83.66$ “ “ “ = 70.4° F; required the altitude A. From these figures we have a mean absolute temperature of $460^\circ + \frac{(77.6 + 70.4)}{2} = 554^\circ$ Abs. Fahr.; hence, from (5), $$A = 26913 \times \frac{30.05}{33.66} \times 2.30258 \times \log_{10} \left( \frac{30.05}{33.66} \right) = 6787.9 \text{ ft}. $$ (Mt. Guanaxato, in Mexico, by Baron von Humboldt.) Strictly, we should take into account the latitude of the place, since $\gamma$, varies with $\varphi$ (see § 76), and also the decrease in the intensity of gravitation as we proceed farther from the earth's centre, for the mercury in the barometer weighs less per cubic inch at the upper station than at the lower. Tables for use in barometric levelling can be found in Trautwine's Pocket-book, and in Seares' Field-book for Railroad Engineers, as also tables of boiling-points of water under different atmospheric pressures, forming the basis of another method of determining heights. 478. Adiabatic Change—Folens' Law.—By an adiabatic change of state, on the part of a gas, is meant a compression or expansion in which work is done upon the gas (in compress- 692 MECHANICS OF ENGINEERING. ing it) or by the gas (in expanding against a resistance) when there is no transmission of heat between the gas and enclosing vessel, or surrounding object, by conduction or radiation. This occurs when the volume changes in a vessel of non-conducting material, or when the compression or expansion takes place so quickly that there is no time for transmission of heat to or from the gas. The experimental facts are, that if a mass of gas in a cylinder be suddenly compressed to a smaller volume its temperature is raised, and its tension increased more than the change of volume would indicate. For example, if a mass of gas at high tension is allowed to expand in a cylinder and drive a piston against a resistance, its temperature falls, and its tension diminishes more rapidly than by Mariotte's law. Again (see Example 3, § 473), if $\frac{1}{V}\frac{dV}{dt}$ of the gas in a rigid vessel, originally at 4 atmos. tension and temperature of 15° Cent., is allowed to escape suddenly through a stop-cock into the outer air, the remainder, while increasing its volume in the ratio $100 : 73$, is found to have cooled to — 27° Cent., and its tension to fall to 28° atmos. The cooling is due, by Mariotte's law, if the temperature had been kept at 28° A. Cent., the tension would have been lowered to $\frac{1}{V}\frac{dV}{dt}$ of 4, i.e., to 2.92 atmospheres only. The reason for this cooling during sudden expansion is, according to the Kinetic Theory of Gases, that since the "sensible heat" (i.e., that perceived by the thermometer), or "hotness" of a gas depends on the velocity of its incessantly moving molecules, and that each molecule after impact with a receding piston has less velocity than before impact, its temperature necesarily falls; and vice versa, when an advancing piston compresses the gas into a smaller volume. If, however, a mass of gas expands without doing work, as when in a vessel of two chambers, one a vacuum, the other full of gas, communication is opened between them, and the gas allowed to fill both chambers, no cooling is noted in the mass as a whole (though parts may have been cooled temporarily). By experiments similar to that in Example 3, § 473, it has ADIABATIC CHANGE—EXAMPLES. 633 been found that for air and the "perfect gases," in an adiabatic change of volume [and therefore of heaviness], the tension varies inversely with the 14th power of the volume. This is called *Poison's Law*. For ordinary purposes (as Weibach suggests) we may use $p$ instead of $1.41$, and hence write $$\text{Adiabat.} \quad p = \left(\frac{v}{v_0}\right)^{-1}, \quad \text{or} \quad p = \left(\frac{V}{V_0}\right)^{-1}; \quad \ldots \quad (1)$$ and combining this relation with the general eqs. (10) and (13), § 472, we have also $$\text{Adiabat.} \quad p = \left(\frac{T}{T_0}\right)^{-1}, \quad \text{or} \quad p = \left(\frac{V}{V_0}\right)^{-1}; \quad \ldots \quad (2)$$ i.e., the tension varies directly as the cube of the absolute temperature; also, $$\text{Adiabat.} \quad p = \left(\frac{T}{T_0}\right)^{-1}, \quad \text{or} \quad p = \left(\frac{V}{V_0}\right)^{-1}; \quad \ldots \quad (3)$$ i.e., the volume is inversely, and the heaviness directly, as the square of the absolute temperature. Here $m$ and $n$ refer to any two adiabatically related states. $T$ is the *absolute temperature*. **Example 1.—Air in a cylinder at 20° Cent. is suddenly compressed to $\frac{1}{4}$ its original volume (and therefore is six times as dense, i.e., has six times the heaviness, as before). To what temperature is it heated? Let $m$ be the initial state, and $n$ the final. From eq. (3) we have** $$\frac{T_n}{T_m} = \sqrt[6]{\frac{6}{1}}; \therefore T_n = 718^\circ\mathrm{Abs.\ Cent.,}$$ or nearly double the absolute temperature of boiling water. **Example 2.—After the air in Example 1 has been given time to cool again to 30° Cent. (temperature of surrounding objects) it is allowed to resume, suddenly, its first volume, i.e.,** A diagram showing a cylinder with a piston moving up and down. 624 MECHANICS OF ENGINEERING. to increase its volume sixfold by expanding behind a piston. To what temperature has it cooled? Here $T_m = 293^\circ$ Abe. Cent., the ratio $V_m : V_a = 4$, and $T_a$ is required. Hence, from (3), $$\frac{T}{293} = \sqrt{\frac{1}{6}}; \therefore T_a = 293 + \sqrt{6} = 119.5^\circ$ Abe. Cent.,$$ or $= -154^\circ$ Cent., indicating extreme cold. From these two examples the principle of one kind of ice-making apparatus is very evident. As to the work necessary to compress the air in Example 1, see § 483. It is also evident why motors using compressed air expansively have to encounter the difficulty of frozen watery vapor (present in the air to some extent). **EXAMPLE 8.** What is the tension of the air in Example 1 (suddenly compressed to its original volume) immediately after the compression, if the original tension was one atmosphere? That is, with $V_a : V_m = 1 : 6$, and $p_a = 14.7$ lbs. per sq. inch, $p_m = ?$ From eq. (1), $(m, lb., sec.)$, $$p_m = 14.7 \times 6^4 = 14.7 \times \frac{216}{6} = 216$$ lbs. per sq. inch; whereas, if, after compression and without change of volume, it cools again to $90^\circ$ Cent., the tension is only $14.7 \times 6 = 88.2$ lbs. per sq. inch (now using Mariotte's law). **47. Work of Expanding Steam following Mariotte's Law.—** Although gases do not in general follow Mariotte's law in expanding (except when heated without special provision for supplying heat), it is found that when saturated steam (i.e., saturated at the beginning of the expansion) in a steam-engine cylinder, when left to expand after the piston has passed the point of "cut-off," diminishes very nearly in accordance with Mariotte's law, which may therefore be applied in this case to find the work done per stroke, and thence the power. In Fig. 527 a horizontal steam-cylinder is EXPANDING STEAM. shown in which the piston is making its left-to-right stroke. The "back-pressure" is constant and $F_0$, $F$ being the area of the piston, and $q$ the intensity (i.e., per unit area) of the back or exhaust pressure on the right side of the piston; while the forward pressure on the left face of the piston = $Fp$, in which $p$ is the steam-pressure per unit area, and is different at different points of the stroke. When the piston moves from $O^{\prime}$ to $D^{\prime}$, $p$ is constant, being $= p_0$ the boiler-pressure, since the steam-port is still open. Between $D^{\prime}$ and $C^{\prime}$ however, the steam being cut off (i.e., the steam-port is closed) at $D^{\prime}$, a distance $a$ from $O^{\prime}$, $p$ decreases with Mariotte's law (nearly), and its value is $(P_0 - F_0)p$, at any point on $C^{\prime}D^{\prime}$, $x$ being the distance of the point from $O^{\prime}$. Above the cylinder, conceive to be drawn a diagram in which an axis $OX$ is parallel to cylinder-axis, $OY$ an axis 7 to the same, while $OZ$ vertically above the left-hand end of the cylinder. As the piston moves, let the value of $p$ corresponding to each of its positions be laid off, to scale, in the vertical immediately above the piston, as an ordinate from the axis $X$. Make $OD^{\prime} = q$ by the same scale, and draw the horizontal line $D^{\prime}C^{\prime}$. Then the effective work done on the piston-rod while it moves through any small distance $dx$ is $$dW = \text{force} \times \text{distance} = F(p - q)dx,$$ and is proportional to the area of the strip RS, whose width is $dx$ and length = $p - q$; so that the effective work of one stroke is $$\int_{O^{\prime}}^{C^{\prime}} dW = F \int_{x=0}^{x=q} (p - q) dx.$$ (1) A diagram showing a piston moving within a cylinder. The x-axis represents time or position along the stroke. The y-axis represents pressure. The curve shows how pressure changes over time. At O', D', C', and D' respectively, pressures are P0, P0 - F0p, P0 - 2F0p, and P0 - 3F0p. A horizontal line represents a constant pressure q. 62 626 MECHANICS OF ENGINEERING. and is represented graphically by the area $A'ARB'C'D'A'$. From $O'$ to $D''p$ is constant and = $p_0$ (while $q$ is constant at all points), and $x$ varies from 0 to $a$; :. $\left[\frac{W}{d'}=F(p_0-q)dx=F(p_0-q)a,\quad\ldots(8)\right.$ which may be called the work of entrance, and is represented by the area of the rectangle $ADD'$. From $D''$ to $C''p$ is variable and, by Mariotte's law, = $\frac{a}{x}p_0$; :. $\left[\frac{W}{d''}=F\int_{D''}(p-q)dx=F\left[ap_0\int_{D''}\frac{ds}{x}-q\int_{D''}ds\right]\right.$ Le., $\left[\frac{W}{d''}=F\left[ap_0\log_a(\frac{l}{a})-q(l-a)\right]\right.\quad\ldots(3)$ = the work of expansion, adding which to that of entrance, we have for the total effective work of one stroke $W=Fp_0a\left[1+\log_a(\frac{l}{a})\right]-Fql.\quad\ldots(4)$ By effective work we mean that done upon the piston-rod and thus transmitted to outside machinery. Suppose the engine to be "double-acting"; then at the end of the stroke a communication is made, by means of the proper valves, between the cylinder and the atmosphere, so that the pressure of the engine; and also between the light of this piston and the boiler (that to the condenser now being closed). On the return stroke, therefore, the conditions are the same as in the forward stroke, except that the two sides of the piston have changed places as regards the pressures acting on them, and thus the same amount of effective work is done as before. If $n$ revolutions of the fly-wheel are made per unit of time (two strokes to each revolution), the effective work done per unit of time, i.e., the power of the engine, is $L=2nW=2nF\left[ap_0\left[1+\log_a(\frac{l}{a})\right]-ql\right].\quad(5)$ WORK OF STEAM-ENGINES. 627 For simplicity the above theory has omitted the consideration of "clearance," that is, the fact that at the point of "cut-off" the mass of steam which is to expand occupies not only the cylindrical volume $F_a$, but also the "clearance" or small space in the steam-passages between the valve and the entrances of the cylinder, the space between piston and valve which is never occupied by steam, and the space between piston and cylinder which is also been disregarded, i.e., the fact that during communication with the boiler the steam-pressure on the piston is a little less than boiler-pressure. For these the student should consult special works, and also for the consideration of water mixed with the steam, etc. Again, a strict analysis should take into account the difference in areas which receive fluid-pressure on the two sides of the piston. EXAMPLE 1.—A reciprocating steam-engine makes 180 revolutions per minute. The steam-pressure is 40 lbs. by the gauge (i.e., $p = 40 + 14.7 = 54.7$ lbs. per sq. inch), the piston area is $F = 120$ sq. in., the length of stroke $l = 16$ in., the steam being "cut off" at $\frac{1}{2}$ stroke ($:a = 4$ in., and $t : a = 4:00$), and the exhaust pressure corresponds to a "vacuum of 25 inches" (by which is meant that the pressure of the exhaust steam will balance 5 inches of mercury), whence $q = \frac{F}{t} = \frac{120}{16} \times 4 = 3.75$ lbs. per sq. inch. Required the work per stroke, $W$, and the corresponding power $L$. Since $t : a = 4$, we have log. $t = 3.309 \times .60906 = 1.386$, and from eq. (4), $(\frac{1}{2} \cdot F \cdot p \cdot l \cdot sec.)$ $$W = \frac{1}{2} \left(54.7 \times 144 \right) \cdot \frac{1}{2} \left(2.386 - \frac{1}{2} \left(2.45 \times 144\right)\right)$$ $$= 5165.86 - 392.0 = 4773.88 \text{ ft. lbs. of work per stroke},$$ and therefore the power at $2$ rev. per sec. (eq. $5$) $$L = 3 \times 9 \times 4773.87 = 19095.5 \text{ ft. lbs. per second}.$$ Hence in horse-powers, which, in ft.-lb.-sec.system, $= L + 550,$ $$Power = 19095.5 + 550 = 34.7 \text{ H.P}.$$ 638 MECHANICS OF ENGINEERING. second by the above engine with given data; assuming with Weibach that the heaviness of saturated steam at a definite pressure (and a corresponding temperature, § 469) is about $\frac{4}{5}$ of that of air at the same pressure and temperature. The heaviness of air at 54.7 lb. per sq. in. tension and temperature 207° Fahr. (see table, § 469) would be, from eq. (15) of § 475 (see also § 469), $$\gamma = \frac{T}{T} \cdot \frac{P}{P} = \frac{0807 \times 492}{400 + 287} = 0.198$$ Ibs. per cub. foot, $\frac{4}{5}$ of which is 0.1237 lbs. per cub. ft. Now the volume $V$ of steam, of this heaviness, admitted from the boiler at each stroke is $$V = Pa = \frac{14.7}{0.1237} = 0.2777$$ cub. ft., and therefore the weight of steam used per second is $$4 \times 0.277 \times 0.1237 = 0.1374$$ lbs. Hence, per hour, $0.1374 \times 3600 = 494.8$ lbs. of feed-water are needed for the boiler. If, with this same engine, the steam is used at full boiler pressure throughout the whole stroke, the power will be greater, viz. $2aF(V_p - q) = 3344$ ft. lbs. per sec., but the consumption $p$ of steam will be four times as great; and hence in economy of operation it will be only 0.44 as efficient (nearly). **480. Graphic Representation of any Change of State of a Con- fined Mass of Gas.—The curve of expansion $AB$ in Fig. 527 is an equilateral hyperbola, the axes $X$ and $Y$ being its asymptotes. If compressed air were used instead of steam its ex- ansion curve would also be an equilateral hyperbola if its temperature could be kept from falling during the expansion (by insuring that heat was supplied), and thus following Mariotte's law would have been possible ($§ 475$). $p$ is con- stant, i.e., $pFx = constant$, and therefore $px = constant$, which is the equation of a hyperbola $p$ being the ordinate and $x$ the abscissa. This curve (dealing with a perfect gas) is also called an **isothermal**, the $x$ and $y$ co-ordinates of its points being pro- *We here neglect the practical fact that a portion of the fresh steam entering the cylinder is condensed prematurely, so that the actual con- sumption is somewhat greater than as here derived.* GRAPHICS OF CHANGE OF STATE OF GAS. 629 portional to the volume and tension, respectively, of a mass of air (or perfect gas) whose temperature is maintained constant. Hence, in general, if a mass of gas be confined in a rigid cylinder of cross-section $P$ (area), provided with an airight piston, Fig. 538, its volume $V$ is proportional to the distance $OD = x$ (of the piston from the closed end of the cylinder) taken as an abscissa, while its tension $p$ at the same instant may be laid off on an ordinate $OZ = D$. Fig. 538. Thus consider $A$ is fixed. Describe an equilateral hyperbola through $A$, asymptotic to $X$ and $Y$, and mark it with the observed temperature (absolute) of the air at this instant. In a similar way the diagram can be filled up with a great number of equilateral hyperbolas, or *isothermal curves*, each for its own temperature. Any point whatever (i.e., above the critical temperature) in the plane angular space $XYZ$ will indicate by its co-ordinates a volume and tension, while the corresponding absolute temperature $T$ will be shown by the hyperbola passing through the point, since these three variables always satisfy the relation $$\frac{PV}{T} = \text{const.}; \text{i.e., } \frac{pFz}{T} = \frac{p_{0}V}{T_{0}}.$$ (1') Any change of state of the gas in the cylinder may now be represented by a line in the diagram connecting the two points corresponding to its initial and final states. Thus, a point moving along the line $AB$, a portion of the isothermal marked 293° Abs. Cent., represents a motion of the piston from $D$ to $E$, and a consequent increase of volume accompanied by just sufficient decrease of pressure that the gas (from its new state) may maintain its temperature at that figure (viz., its temperature at 630 MECHANICS OF ENGINEERING. A). If the piston move from $D$ to $E$, without transmission of heat, i.e., adiabatically (§ 478), the tension falls more rapidly, and a point moving along the line $AB$ represents the corresponding continuous change of state. $AB$ is a portion of an adiabatic curve, whose equation, from § 478, is $$\frac{P}{P_2} = \left(\frac{F_{ax}}{F_{ax'}}\right), \quad \text{or} \quad Pd = p_2e^x = \text{const}. \quad (1)$$ in which $p_x$ and $x$ refer to the point $K$ where this particular adiabatic curve cuts the isothermal of freezing-point. Evidently an adiabatic may be passed through any point of the diagram. The mass of gas in the cylinder may change its state from $A$ to $B$ by an infinite number of routes, or lines on the diagram, all differing only in their rate of change. This is likely to occur for a rapid motion of the piston. For example, we may cool it without allowing the piston to move (and hence without altering its volume nor the abscissa $x$) until the pressure falls to a value $p_x = DL = EB$, and this change is represented by the vertical path from $A$ to $L$; and then allow it to expand, and push the piston from $D$ to $E$ (i.e., do external work), during which expansion heat is to be supplied at just such a rate as to keep the constant, $p_x = p_2$, this latter change corresponding to the horizontal path $LB$ from $L$ to $B$. It is further noticeable that the work done by the expanding gas upon the near face of the piston (or done upon the gas when compressed) when the space $dx$ is described by the piston, is $$= Fpdx,$$ and therefore is proportional to the area $\rho dx$ of the small vertical strip lying between the axis $X$ and the line or route showing the change of state; whence the total work done on the near piston-face, being $$= \int Fpdx,$$ represented by the area $\rho dx$, is equal to that between the initial and final ordinates, the axis $X$ and the particular route followed between the initial and final states (N.B. We take no account here of the pressure on the other side of the piston, the latter depending on the style of engine). For example, the work done on the near face of the piston during adiabatic expansion WORK OF ADIABATIC EXPANSION. 631 from $D$ to $E$ is represented by the plane figure $A'B'ED$, and is measured by its area. The mathematical relations between the quantities of heat imparted or rejected by conduction and radiation, and transformed into work, in the various changes of which the confined gas is capable belong to the subject of Thermodynamics, which cannot be entered upon here. It is now evident how the cycle of changes which a definite mass of air or gas experiences when used in a hot-air engine, compressed-air engine, or air-compressor, is rendered more intelligible by the aid of such a diagram as Fig. 529; but it must be remembered that during the entrance into, or exit from, the cylinder, of the mass of gas used in one stroke, the distance $x$ does not represent its volume, and hence the locus of the points in the diagram determined by the co-ordinates $p$ and $x$ (Fig. 529) will not do so. This is due to the fact that in this state in way just explained for the mass when confined in the cylinder. However, the work done by or upon the gas during entrance and exit will still be represented by the plane figure included by that locus (usually a straight horizontal line, pressure constant) and the axis of $X$ and the terminal ordinates. 481. Adiabatic Expansion in an Engine using Compressed Air. ---Fig. 529. Let the compressed air at a tension $p_m$ and an absolute temperature $T_m$ be supplied from a reservoir (in which the loss is continuously compensated by a fresh compressor). Neglecting the resistance of the port, its tension and temperature when behind the piston are still $p_m$ and $T_m$. Let $x_0 =$ length of stroke, and let the cut-off (or closing of communication with the reservoir) be made at some point $D$ where $x = x_0$, the position of which being so chosen (i.e., the ratio $\frac{x_0}{x}$), so convenient that adiabatic expansion from $D$ to $E$ the pressure shall have fallen from $p_m$ at $M$ (state in) to a value $p_a = p_d$ A diagram showing a compression process in an engine. The x-axis represents distance \( x \), and the y-axis represents pressure \( p \). The curve shows how pressure decreases as distance increases. The points on the curve are labeled with their respective values. Fig. 529 632 MECHANICS OF ENGINEERING. = one atmosphere at $N$ (state $n$), at the end of stroke ; so that when the piston returns the air will be expelled (" exhausted ") at a tension equal to that of the external atmosphere (though at a low temperature). Hence the back-pressure at all points either way will be = $p_a$, per unit area of piston, and hence the total back-pressure = $Fp_a$, $F$ being the piston area. From $O$ to $D$ the forward pressure is constant and = $Fp_a$, and the effective work, therefore, or work on piston-rod from $O$ to $D_1$ is Work of entrance = $\int_{O}^{D_1} W = F(p_m - p_s)z_m$. . . . (1) represented by the rectangle $MMLN$: The out-off being made at $D$, the volume of gas now in the cylinder, viz., $V_m = Fz_m$, is left to expand. Assuming no device adopted (such as injecting hot-water spray) for preventing the cooling and rapid decrease of tension during expansion, the latter is adiabatic, and hence the tension at any point $P$ between $M$ and $N$ will be $p = p_a \left(\frac{z_m}{z}\right)^{\frac{1}{\gamma}}$. . . [see § 478; $V = Fx$]; . . . (a) ∴ Work of expansion $\int_{D}^{D_1} W = \int_{z_m}^{z_0} F(p - p_s)dz = F \int_{z_m}^{z_0} pdz = Fp_a(z_m - z_n)$, (2) and is represented by the area $MPNL$. But $\int_{z_m}^{z_0} pdz = p_a z_m \int_{z_m}^{z_0} z^{-1} dz = -2p_a z_m \left[\left(\frac{1}{z_m}\right)^{\frac{1}{\gamma}} - \left(\frac{1}{z_0}\right)^{\frac{1}{\gamma}}\right]$; i.e., $F \int_{z_m}^{z_0} pdz = 2Fp_a z_m \left[1 - \left(\frac{z_m}{z_0}\right)^{\frac{1}{\gamma}}\right]$. . . (3) Substitute (3) in (2) and then add (2) to (1), noting that **COMPRESSED-AIR ENGINE** 633 $$F(p_a - p_s)\omega_m = Fp_s(\alpha_s - \alpha_a) = Fp_s\omega_m\left[1 - \frac{z_aP_a}{z_aP_s}\right],$$ which furthermore, since $n$ and $m$ are adiabatically related (see $(a)$), can be reduced to $$Fp_s\omega_m\left[1 - \left(\frac{z_a}{z_s}\right)^{\gamma}\right],$$ and we have finally : Total work on piston-rod per stroke $\left\} = W = 3Fx_mp_s\left[1 - \left(\frac{z_a}{z_s}\right)^{\gamma}\right]. \quad (4)$ But $Fx_m = V_m$, and the adiabatic relation holds good, $$\left(\frac{V_a}{V_s}\right) = \left(\frac{P_a}{P_s}\right)^{\gamma}; \quad \text{i.e.,} \quad \left(\frac{z_a}{z_s}\right)^{\gamma} = \left(\frac{P_a}{P_s}\right);$$ therefore we may also write $$W = 3V_mp_s\left[1 - \left(\frac{P_a}{P_s}\right)^{\gamma}\right]; \quad \ldots \ldots (5)$$ in which $V_m =$ the volume which the mass of air used per stroke occupies in the state $m$, i.e., in the reservoir, where the tension is $p_a$, and the absolute temperature is $T_a$. To find the work done per pound of air used (or other unit of weight), we must divide $W$ by the weight $G = V_mp_s$ of the air used per stroke, remembering (eq. (13), § 472) that $$V_mp_s = [V_mp_sT_a] + (T_ap_s).$$ Work per unit of weight of air used in adiabatic working $\left\} = 3T_a\frac{P_a}{T_{a}}\left[1 - \left(\frac{P_a}{P_s}\right)^{\gamma}\right]. \quad (6)$$ In $(6)$, $y_s = 0807$ lbs. per cub. foot, $p_s = 14.7$ lbs. per sq. inch, and $T = 273^{\circ}$ Abs. Cent. or $492^{\circ}$ Abs. Fahr. 634 MECHANICS OF ENGINEERING. It is noticeable in (6) that for given tensions $p_m$ and $p_a$, the work per unit of weight of air used is proportional to the absolute temperature $T_m$ of the reservoir. The temperature $T_m$ to which the air has cooled at the end of the stroke is obtained as in Example 2, § 478, and may be far below freezing point unless the reservoir is very large. EXAMPLE. Let the cylinder of a compressed-air engine have a section of $F = 108$ sq. in. and a stroke $x_m = 15$ inches. The compressed air entering the cylinder is at a tension of 3 atmos. (i.e., $p_m = 29.4$ lbs. per sq. in., and $p_a + p_m = 4$), and at a temperature of 97° Cent. (i.e., $T_m = 300$ Abs. Cent.). Required the proper point of cut-off, or $x_a = t$, in order that the tension may fall to one atmosphere at the end of the stroke; also the work per stroke, and the work per pound of air. Use the foot, pound, and second. From eq. (a), above, we have $$x_m = x_a \left( \frac{P_m}{p_a} \right) = 1.25 \sqrt{\frac{T}{4}} = 0.7875 \text{ ft} = 9.45 \text{ inches},$$ and hence the volume of air in state m, used per stroke [eq. (5)] is $$V_m = Fx_m = 144 \times 0.7875 = 0.5906 \text{ cubic feet};$$ while the work per stroke is $$W = 3 \times 0.5906 \times 29.4 \times 144 \times [1 - (\frac{1}{2})^2] = 1545 \text{ ft}.\text{ lb.},$$ and the work obtained from each pound of air, eq. (6), $$= 3 \times 300 \times \frac{14.7 \times 144}{0.0807 \times 273} \times [1 - \sqrt[3]{\frac{T}{2}}] = 17810$$ ft. lb. per pound of air used. The temperature to which the air has cooled at the end of stroke [eq. (2), § 478] is **COMPRESSED-AIR ENGINE.** $$T_a = T_n \sqrt{\frac{P_2}{P_0}} = 300 \times \sqrt{\frac{754}{300}} = 300 \times .754 = 238^\circ \text{Abe. C.}$$ I.e., - 35° Centigrade. **482. Remarks on the Preceding.—This low temperature is objectionable, causing, as it does, the formation and gradual accumulation of snow, from the water vapor usually found in small quantities in the air, and ultimate blocking of the ports, owing to high pressure to $T_a$ (see fig. 6). The temperature of the reservoir, $T_r$, will be correspondingly higher, and also the work per pound of air, eq. (6). If the cylinder be encased in a “jacket” of hot water, or if spray of hot water be injected behind the piston during expansion, the temperature may be maintained nearly constant, in which event Mariotte's law will hold for the expansion, and more work will be obtained per pound of air; but the point of cut-off must be differently placed. Thus if, by eq. (4), § 479, we make the back-pressure, which is $P_b$, $Fb$ times to $T_b$, then at the end of the stroke by Mariotte's law, we have $$\text{Work per stroke with } \left\{ \right\} = Fap_b \log_{a} \left( \frac{T}{T_b} \right) = V_s p_b \log_{a} \left( \frac{T}{T_b} \right),$$ and hence $$\text{Work per unit of weight of air}, \left\{ \right\} = T_a \frac{P}{y_r T_s} \log_{a} \left( \frac{T}{T_b} \right).$$ Applying these equations to the data of the example, we obtain $$\text{Work per unit of weight of air with iso-thermal expansion}, \left\{ \right\} = 0.69 T_a \frac{P_2}{y_r T_s},$$ whereas, with adiabatic expansion, work per unit of weight of air is only $$\left\{ \right\} = 0.69 T_a \frac{p_2}{y_r T_s}.$$ 636 MECHANICS OF ENGINEERING. 483. Double-acting Air-compressor, with Adiabatic Compression. This is the converse of § 431. In Fig. 530 we have the piston moving from right to left, compressing a mass of air which at the beginning of the stroke fills the cylinder. This is brought about by means of an external mechanical force (e.g., a screw-piston, etc.) which exerts a thrust or pull along the piston-rod, enabling it with the help of the atmospheric pressure of the fresh supply of air flowing in behind it, to first compress a cylinder-full of air to the tension of the compressed air in the reservoir, and then, the port or valve opening at this stage, forms a definite volume of compressed air. Let the temperature and tension of the cylinder-full of fresh air be $T_a$ and $p_m$, and the tension in the reservoir be $p_{m_r}$. Suppose as the piston passes from $E$ toward the left, the air on the left has no escape and is compressed, its tension and temperature increasing adiabatically until it reaches a value $p_{m_t}$ that is in reservoir, at which instant, the piston being at some point $D$, a valve opens and the further progress of the piston is purely due to its own weight. It leaves the re- servoir without further increasing its tension. Throughout the whole stroke the piston-rod has the help of one atmosphere pressure on the right face, since a new supply of air is entering on the right to be compressed in its turn on the return stroke. The work done from $E$ to $D$ may be called the work of compression; that from $D$ to $O$, the work of delivery. [Since, here, $dx$ and $dW$ (or increment of work) have contrary signs, we introduce the negative sign as shown.] The work of compression = $\int_E^D \left( p - p_m \right) dx \dots \dots (1c)$ The work of delivery = $\int_D^O \left( p_m - p_m \right) dx \dots \dots (1d)$ Fig. 530. AIR-COMPRESSOR. 637 In these equations only $p$ and $x$ are variables. In the summation indicated in (10) $p$ changes adiabatically; in (1d) $p$ is constant = $p_m$, as now written. In the adiabatic compression the air passes from the state $m$, to the state $m_1$ (fig. 5, M, in figure). The substitutions in these equations being of the same form as those in equations (1) and (2) of § 481, but with limits inverted, we may write immediately, $$Work\ per\ stroke\ =\ W=3\ V_{m_1}p_{m_1}\left[1-\left(\frac{p_{m_1}}{p_m}\right)^{\gamma}\right]\ .\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (9)$$ and $$Work\ per\ unit\ of\ weight\ =\ 3T_{m_1}\frac{p_{m_1}}{y_{m_1}T_{m_1}}\left[1-\left(\frac{p_{m_1}}{p_m}\right)^{\gamma}\right]\ .\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (3)$$ The value of $T_{m_1}$, at the immediate end of the sudden compression, by eq. (3) of § (478), is $$T_{m_1}=T_{m_1}\left(\frac{p_{m_1}}{p_m}\right)^{\gamma}\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\ .\\ T_{m}=T_{m}\left(\frac{p_m}{p_m}\right)^{\gamma}\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.\ ,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,.,. (4) The temperature of the reservoir being $T_m$, as in § 481 (usually much less than $T_m$), the compressed air entering it cools down gradually to that temperature, $T_m$, contracting in volume from its initial state to its final state, under the same tension $p_m$. The mechanical equivalent of this heat is lost. Let us now inquire what is the efficiency of the combination of air-compressor and compressed-air engine, the former supplying air for the latter, both working adiabatically, assuming that no tension is lost by the compressed air in passing along the reservoir between, i.e. that $p_m=p_n$. Also assume (as already implied, in fact) that $p_n=p_m=\mathrm{one~atmos}$, and that the temperature, $T_n$, of the air entering the compressor cylinder is equal to that, $T_m$, of the reservoir and transmission pipe. To do this we need only find the ratio of the amount of work obtained from one pound (or other unit of weight) in the compressed-air engine to the amount spent in compressing one pound of air in the compressor. Calling this ratio $\eta$, the 638 MECHANICS OF ENGINEERING. efficiency, and dividing eq. (6) of § 481 by eq. (3) of this paragraph, we have, with substitutions just mentioned, $$\eta = \frac{T_{m}}{T_{a}} = \frac{\text{Abs. temp. of outer free air}}{\text{Abs. temp. of air at end of sudden compression}},$$ (5) or, substituting from eq. (4), and remembering that $T_{m} = T_{a}$, we have also $$\eta = \left(\frac{p_{a}}{p_{m}}\right)^{1},$$ (6) also, since $$\left(\frac{p_{a}}{p_{m}}\right)^{1} = \frac{T_{a}}{T_{m}},$$ we may write $$\eta = \frac{T_{a}}{T_{m}} = \frac{\text{Ab. tem. air leaving eng. cyl.}}{\text{Ab. tem. outer free air}},$$ (7) For practical details of the construction and working of engines and compressors, and the actual efficiency realized, the student may consult special works, as they lie somewhat beyond the scope of the present work. **EXAMPLE 1.—** In the example of § 445, the ratio of $p_m$ to $p_a$ was $\frac{1}{4}$. Hence, if compressed air is supplied to the reservoir under above conditions, the efficiency of the system is, from eq. (6), $\eta = F_2/F_1 = 0.784$, about 80 per cent. **EXAMPLE 2.—** If the ratio of the tensions is as small as $\frac{F_2}{F_1} = \frac{1}{6}$, the efficiency would be only ($\eta$) = 0.55; i.e., 45 per cent of the energy spent in the compressor is lost in heat. **EXAMPLE 3.—** What horse-power is required in a blowing engine to furnish 10 lbs. of air per minute at a pressure of 4 atmos., with adiabatic compression, the air being received by the compressor at one atmosphere tension and 27 Cent. (f.t.b.sec. system). Since $27^{\circ}\mathrm{C} = 300^{\circ}\mathrm{C}$. Abs. C. = $T_m$, we have, from eq. (4), $$T_m = 300 (\text{ft})^{1} = 477^{\circ}\text{ Abs. Cent.};$$ and hence, eq. (3), A diagram showing a simple mechanical system with a piston moving up and down in a cylinder. HOT-AIR ENGINES. 630 The work per pound of air = $8 \times 477 \times 144 \left[1 - \left(\frac{1}{4}\right)^{\frac{1}{4}}\right]$ = 80870 ft. lb. per pound of air. Hence 10 lb. of air will require 508700 ft. lb. of work; and if this is done every minute we have the reg. H. P. = $\frac{508700}{60} = 8478$ ft. lb. per min. Norr.—If the compression could be made isothermal, an approximation to which is obtained by injecting a spray of cold water, we would have, from eqs. (1) and (2) of § 482: $$W_{per lb. air} = T_{air} \cdot \frac{P_{a}}{T_{air}} \log \left(\frac{p_{m}}{p_{n}}\right) = 300 \times 14.7 \times 144 \times 0.087 \times 273 \times 1.386$$ = 39950 ft. lb. per lb., and the corresponding H. P. = 19.1; a saving of about 35 per cent, compared with the former. The difference was employed in heating the air in the air-compressor with adiabatic compression, and was lost when extra heat was dissipated in the reservoir as the air cooled again. This difference is easily shown graphically by comparing in the same diagram the areas representing the work done in the two cases. 484. Hot-air Engines.—Since we have seen that the tension of air and other gases can be increased by heating, if the volume be kept the same, a mass of air thus treated can afterwards be allowed to expand in a working cylinder, and thus become a means of converting heat into work. In Stirling's hot-air engine a definite confined mass of air is used indefinitely without loss (except that occasional small supplies are needed to make up for leakage), and is alternately heated and cooled. A displacer piston, or plunger, is placed in a hollow chamber, so connected with the piston of the working cylinder and the fly-wheel, that its forward stroke is made while the other piston waits at the beginning of its stroke. In this motion the plunger causes the confined air to pass in a thin sheet over the top and sides of the furnace dome, thus greatly increasing its tension. The air then expands behind the working piston with falling tension and temperature, and, A diagram showing a hot-air engine with a plunger and flywheel. 640 MECHANICS OF ENGINEERING. while that piston pauses at the end of its forward stroke, is again shifted in position, though without change of volume, by the return stroke of the plunger, in such a way as to pass through the space between the two cylinders, and so to mix some of its heat, which it re-absorbs to some extent when, after further cooling in the "refrigerator" or pipe coil and compression by the return stroke of the piston, it is made to pass backward through the regenerator to be further heated by the furnace in readiness for a forward stroke. This feature, however, has not realized all the expectations of its inventor and improver, as to economy of heat and fuel. In *Eriocson's* hot-air engine, of more recent date, the displacement of air is effected by a plunger, but valves can be opened through its edges when moving in one direction, thus causing it to act temporarily as a loose plunger, or shifter. The two pistons move simultaneously in the same direction in the same cylinder, but through different lengths of stroke, so that the space between them is alternately enlarged and contracted. The working piston also has valves opening through it for receiving a fresh supply of air into the space between the two pistons. During the forward stroke of the first instalment of air is entering through proper valves into the space between it and the other, whose valves are now closed and which is now expelling from its further face, through proper valves, the air used in preceding stroke; no work is done in this stroke. On the return stroke this fresh supply of air is free to expand behind the now retracting working piston, while its tension is greatly increased by being shifted (at least a large portion of it) over the furnace GAS-ENGINES. 641 dome through the valves (now open) of the plunger piston, by the motion of the latter, which now acts as a loose plunger. The engine is therefore only single-acting, no work being done in each forward stroke. For further details, see Goodwin's and Rankine's works on the steam-engine; also the article "Hot-air Engine" in Johnson's Cyclopaedia by Pres. Barnard, and Röntgen's Thermodynamics. 485. Brayton's Petroleum-engine.--Although a more recent invention than the gas-engines to be mentioned presently, this motor is more closely related to hot-air engines than the latter. By a slow combustion of petroleum vapor the gaseous products of combustion, while under considerable tension, are enabled to follow up a piston with a sustained pressure, being left to expand through the latter part of the stroke. Thus we have the furnace and working cylinder combined in one. The gradual expansion of the gases, according to the analogy of the principle of the Davy safety-lamp that flame will not spread through layers of wire guaze of proper fineness. 486. Gas-engines.--We again have the furnace and working cylinder in one in a "gas-engine," where illuminating gas and atmospheric air are introduced into the working cylinder in proper proportions (about ten parts of air to one of gas, by weight) so that when the mixture is ignited it does not explode at a certain point of the stroke, causing a very sudden rise of temperature and tension, after which the mass expands behind the piston with falling pressure. On the return stroke the products of combustion are expelled, and no work done, these engines being single-acting. In some forms the mixture is compressed before explosion, since it has been found that under this treatment a mixture containing a larger proportion of air to gas can be made to ignite, and that then the resulting pressure is greater than when ignited without compression or of the mixture in the Brayton engine. That is, the effect is analogous to that of "slow-burning powder" in a gun. In the "Otto Silent Gas-engine" the explosion occurs only every fourth stroke, and one side of the piston is always open 643 MECHANICS OF ENGINEERING. to the air. The action on the other side of the piston is as follows: (1) In the forward stroke a fresh supply of explosive mixture is drawn into the cylinder at one atmosphere tension. (2) The next (backward) stroke compresses the mixture into about one fourth of its original bulk, this operation occurring at the expense of the kinetic energy of the fly-wheel. (3) The mixture is ignited, the pressure rises to 8 or 10 atmospheres, and work is done on the piston through the next (forward) stroke, the tension of the products of combustion having fallen to about one atmosphere at the end of the stroke. (4) In the last (backward) stroke the products of combustion are expelled and no work is done. The Atkinson “Cycle Gas-engine,” an English invention of recent date (see the London Engineer for May 1887 pp. 361 and 380) also makes an explosion every fourth stroke, but the link work connecting the piston and fly-wheel is of such design that the latter makes but one revolution during the four strokes. Also the length of the expansion or working stroke is greater than that of compression stroke and the products of combustion are more completely decomposed. Consequently, the efficiency of this motor is at present greater than that of any other gas-engine. See § 487. One of the most simple gas-engines is made by the Economic Motor Company of New York. The piston has no packing, being a long plunger ground to fit the cylinder accurately and kept well lubricated. As with most gas-engines the cylinder is encased in a water-jacket to prevent excessive heat from affecting the working parts. Consequently decomposition of fuel is insignificant. For further details on these engines see Bockris's "Steam engine," Clark's "Gas-engine" in Van Nostrand's Science Series, and article "Gas-engine" in Johnson's Cyclopaedia; also Prof. Thurston's report on Mechanical Engineering at the Vienna Exhibition of 1873, and proceedings of the "Society of Engineers" (England) for 1881. 487. Efficiency of Heat-engines.—According to the mechanical theory of heat, the combustion of one pound of coal, pre- GAS-ENGINES. 645 DUCING, AS IT DOES, ABOUT 14,000 HEAT-UNITS (BRITISH THERMAL UNITS; SEE § 148, MECHANICS) SHOULD FURNISH $$14,000 \times 772 = 10,808,000 \text{ ft. lbs. of work},$$ IF ENTIRELY CONVERTED INTO WORK. LET US SEE HOW NEARLY THIS IS ACCOMPLISHED IN THE PERFORMANCE OF THE MOST RECENT AND ECONONIMAL MARINE ENGINES OF THE PRESENT DAY, VIZ., THE TRIPLE EXPANSION ENGINES OF SOME ATLANTIC STEAMERS, WHICH ARE CLAIMED TO HAVE CONSUMED PER HOUR ONLY 1.25 LBS. OF COAL FOR EACH MEASURED ("INDIRECT") HORSEPOWER OF EFFECTIVE WORK DONE IN THEIR CYLINDERS. THE WORK-EQUIVALENT OF 1.25 LBS. OF COAL PER HOUR IS $$1.25 \times 14,000 \times 772 = 13,510,000 \text{ ft. lbs. per hour};$$ WHILE THE ACTUAL WORK PER HOUR IMPLIED IN "ONE H.P. PER HOUR" IS $$33000 \times 60 = 1,980,000 \text{ ft. lbs. per hour}.$$ THAT IS, THE ENGINE UTILIZES ONLY ONE SEVENTH OF THE HEAT OF COMBUSTION OF THE FUEL. ACCORDING TO PROF. THURSTON, THIS IS A RATHER EXTRAVAGANT CLAIM (1.35), THE ACTUAL CONSUMPTION HAVING PROBABLY BEEN 1.4 LBS. OF COAL PER HOUR. THE ORDINARY COMPOUND MARINE ENGINE IS STATED TO USE AS LITTLE AS 2.00 LBS. PER HOUR FOR EACH H.P. MOST OF THE HEAT NOT UTILIZED IS DISSIPATED IN THE CONDENSER. SIMILARLY, THE WATER-JACKET, A NECESSARY EVIL IN THE OPERATION OF THE GAS-ENGINE, IS A SOURCE OF GREAT LOSS OF HEAT AND WORK. STILL, MR. WM. ANDERSON IN HIS RECENT WORK, "CONVERSION OF HEAT INTO WORK" (LONDON, 1887), MENTIONED A MOTOR OF THIS CLASS AS HAVING CONVERTED INTO WORK $\frac{4}{7}$ OF THE HEAT OF COMBUSTION [IAN OWEN'S "STOKES' GAS ENGINE"] AT PHILADELPHIA BY THE SIEGMANN INSTITUTIONS (HOBOKEN, N.J., 1883); WHILE PROF. UNEV FOUND THE ATKINSON ENGINE (SEE LAST PARAGRAPH) CAPABLE OF RETURNING (IN THE CYLINDER) FULLY $\frac{4}{7}$ OF THE HEAT-EQUIVALENT OF THE GAS CONSUMED. [THIS LATTER RESULT WAS CONFIRMED IN PHILADELPHIA IN JAN. 1889 BY PROF. BARR, UNDER DIRECTION OF PROF. THURSTON.] A page from a book with text discussing gas engines and their efficiency. 644 MECHANICS OF ENGINEERING. 488. Duty of Pumping-engines.—Another way (often used in speaking of the performance of pumping-engines) of expressing the degree of economy attained in the use of fuel by the combined furnace, boiler, and engine is to give the number of foot-pounds of work obtained from each 100 lb. of coal consumed in the furnace, calling it the "duty" of the engine. For example, suppose that from each pound of coal 990,000 ft. lbs. of work are obtained. From this we gather that, since one horse-power consists of $33,000 \times 60 = 1,980,000$ ft. lbs. per hour, the engine mentioned must use each hour $$1,980,000 + 990,000 = 2 \text{ lb. of coal for each H. P. developed;}$$ which is as low a figure as that attained by the marine engines last quoted. 489. Buoyant Effort of the Atmosphere.—In the case of a body of large bulk but of small specific gravity the buoyant effort of the air (due to the same cause as that of water, see § 436) becomes quite appreciable, and may sometimes be greater than the weight of the body. This buoyant effort is equal to the weight of air displaced, i.e., $V\gamma$, where $V$ is the volume of air displaced, and $\gamma$ its heaviness. If $G_1 =$ total weight of the body producing the displacement, the resultant vertical force is $$P = G_1 - V\gamma,$$ (1) and for equilibrium, or suspension in the air, we must have $$P = 0,$$ i.e., $$G_1 = V\gamma.$$ (2) We may therefore find approximately the elevation where a given balloon will cease to ascend, by determining the heaviness $\gamma$ of the air at that elevation from eq. (9); then, knowing approximately the temperature of the air at that elevation, we may compute its tension $p$ (eq. (13), § 472), and finally, from eqs. (3), (4), or (5) of § 477, obtain the altitude required. **EXAMPLE**—The car and other solid parts of a balloon weigh BALLOONS. 645 400 lbs., and the bag contains 13,000 cub. feet of illuminating gas weighing 0.030 lbs. per cub. foot at a tension of one atmosphere and temperature of 15° Cent., so that its total weight = 12,000 × 0.030 = 360 lbs. Hence $G_1 = 760$ lbs. We may also write with sufficient accuracy : whole volume of displacement = $V = 12,000$ cub. ft. As the balloon ascends the exterior pressure diminishes, and the confined gas tends to expand and so increase the volume of the balloon. In order to prevent this we shall suppose prevented by the strength of the envelope. At the surface of the ground (station $n$ of Fig. 531 ; see also Fig. 526) let the barometer read 39.6 inches and the temperature be 15° Cent. Then $T_n = 288$ Abs. Cent., and the heaviness of the air at $n$ is $$\gamma_n = \frac{.087 \times 273}{14.7} \times \frac{288}{288} = .087 \times \frac{273}{14.7} = .0754 \text{ lbs. per cub. ft.}$$ $$\left(= \frac{\gamma_n T_n p_n}{p_n T_n}\right) = .087 \times \frac{273}{14.7} = .0754 \text{ lbs. per cub. ft.}$$ At the unknown height $h$, where the balloon is to come to rest, i.e., at $M$, $G_1$ must = $\gamma_M$ [eq. (9)] ; therefore $$\gamma_m = \frac{G_1}{V} = \frac{760}{12,000} = .0633 \text{ lbs. per cub. ft.;}$$ and if the temperature at $M$ be estimated to be 5° Cent. (or $T_m = 278$ Abs. Cent.) (on a calm day the temperature decreases about 1° Cent. for each 500 ft. of ascent), we shall have, from $\frac{p_n}{\gamma_n T_n}$ : $$\frac{p_n}{\gamma_n T_n} = \frac{p_m}{\gamma_m T_m}$$ $$\frac{p_n}{\gamma_n T_n} = \frac{.0754}{288} \times \frac{288}{278} = 1.206;$$ and hence, from eq. (5), § 477, with $\frac{1}{T_m + T_a}$ put for $T_a$, $$h = 26913 \times \frac{1}{288} \times 23058 \times \log_{10} 1.206 = 5088 \text{ ft.}$$ CHAPTER VI HYDRODYNAMICS REGUIN—STeady Flow of LIQUIDS THROUGH PIPES AND ORIFICES. 489. The subject of Water in Motion presents one of the most unsatisfactory branches of Applied Mechanics, from a mathematical stand-point. The internal eddies, cross-currents, and general intricacy of motion of the particles among each other, occurring in a pipe transmitting a fluid, are almost entirely defamant of mathematical expression, though the flow of water through a circular orifice in a thin plane into the air presents a case comparatively simple. The theory of this flow is probably quite close to the truth. In most practical cases we are forced to adopt as a basis for mathematical investigation the simple assumption that the particles move side by side in such a way that those which at any instant form a lamina or thin sheet, $\gamma$ to the axis of the pipe or orifice, remain together as a lamina during the further stages of the flow. This is the Hypothesis of Flow in Plane Layers, or Laminated Flow. Experiments have failed to make apparent the discrepancies between the indications of the formulae resulting from this theory and the actual results of practice; so that the science of Hydrodynamics is largely one of coefficients determined by experiment. 490. Experimental Phenomena of a "Steady Flow"—As preliminary to the analysis on which the formulae of this chapter are based, and to acquire familiarity with the quantities involved, it will be advantageous to study the phenomena of the apparatus represented in Fig. 532. A large tank or reservoir $BC$ is connected with another, $DE$, at a lower level, by means of a rigid pipe opening under the water-level in each tank. This 646 STeady Flow of a Liquid. 647 pipe has no sharp curves or bends, is of various sectional areas at different parts, the changes of section being very gradual, and the highest point $N$, not being more than 30 ft. higher than $BC$, the surface-level of the upper tank. Let both tanks A diagram showing a pipe with a stopcock and a water tank. Fig. 528 be filled with water (or other liquid), which will also rise to $H$ and to $K$ in the pipe. Stop the ends $L$ and $N_1$ of the pipe, and through $M$, a stop-cock in the highest curve, pour in water to fill the remainder of the pipe; then, closing $M$, open up $L$ and $N_1$. If the dimensions are not extreme (and subsequent formulae will furnish the means of testing such points) the water will now begin to flow from the upper tank into the lower, and all parts of the pipe will continue full of water as the flow goes on. Further, suppose the upper tank so large that its surface-level sinks very slowly; or that an influx at $A$ continually makes good the efflux at $E$; then the flow is said to be a Steady Flow; or, a state of permanency is said to exist; i.e., the circumstances of the flow at each section of the pipe are permanent, or steady. 648 MECHANICS OF ENGINEERING. By measuring the volume, $V$, of water discharged at $E$ in a time $t$, we obtain the volume of flow per unit of time, viz., $$Q = \frac{V}{t}.$$ (1) while the weight of flow per unit of time is $$Q = Qy,$$ (2) where $y$ = heaviness ($\S$ 7) of the liquid concerned. Water being incompressible and the pipe rigid, it follows that the same volume of water per unit of time must be passing at each cross-section of the pipe. But this is equal to the volume of a prism of water having $F_y$ the area of the section, as a base, and as an altitude, the mean velocity with which the liquid particles pass through the section. Hence for any section we have $$Q = Fe = constant = F_yv = F_yu,$$ etc. (Equat. of continuity.) (3) in which the subscripts refer to different sections. If the flow were unsteady, e.g., if the level $BC$ were sinking, this would be true for a definite instant of time; but when steady, we see that it is permanently true; e.g., $F_y$, at any instant = $Fe$, at the instant immediately subsequent or previous. In other words, in a steady flow the velocity at a given section remains unchanged with lapse of time. [N.B. We here assume for simplicity that the different particles of water passing simultaneously through a given section (i.e., abreast of each other) have equal velocities, viz., the velocity which all other particles will assume on reaching this section. Strictly, however, the particles at the sides are some- what retarded by friction on the surface of the pipe. This assumption leads to what is called "Parallel Flow," or "Flow in Plane Layers, or Laminated Flow." Let us suppose $Q$ to have been found as already prescribed. We may then, knowing the internal sectional areas at different parts of the pipe, $N_1$, $N_2$, etc., compute the velocities STeady Flow of liquids. 649 $$v_{s} = Q + F_{s}, \quad v_{r} = Q + F_{r}, \quad \text{etc.,}$$ which the water must have in passing those sections, respectively. It is thus seen that the velocity at any section has no direct connection with the height or depth of the section from the plane, $BC$, of the upper reservoir surface. The fraction $$\frac{s'}{2g}$$ will be called the height due to the velocity, $v$, or simply the velocity-head, for convenience. Next, as to the value of the internal fluid pressure, $p_s$ per unit-area (in the water itself and against the side or wall of pipe) at different sections of the pipe. If the end $N_r$ of the pipe were stopped, the problem would be one in Hydrostatics, and the pressure against the side of the pipe at $N_r$ (also at $N_s$ on same level) would be simply $$p_s = p_a + h_s,$$ measured by a water column of height $$\frac{P_s}{Y} = \frac{p_a}{Y} + h_s = b + h_s,$$ in which $p_a$ = one atmosphere, and $b$ = 34 ft. = height of an ideal water barometer, and $Y$ = 62.5 lbs. per cubic foot; and this would be shown experimentally by screwing into the side of the pipe at $N_r$ a small tube open at both ends; the water would rise in it to the level $BC$. That is, a column of water of height $= h_s$, would be sustained in it, which indicates that the internal pressure at $N_r$ corresponds to an ideal water column of a height $$\frac{P_s}{Y} = b + h_s.$$ But when a steady flow is proceeding, the case being now one in Hydrodynamics, we find the column of water sustained at rest in the small tube (called an open piezometer), $N_rS$ has a height $y_r$, less than $h_s$, and hence the internal fluid pressure is 650 MECHANICS OF ENGINEERING. less than it was when there was no flow. This pressure being called $p_1$, the ideal water column measuring it has a height $$\frac{p_1}{y} = b + y \dots \dots \dots (4)$$ at $N_1$, and will be called the pressure-head at the section referred to. We also find experimentally that while the water is steadily the piezometer-height $y$ (and therefore the pressure-head $b + y$) at any section remains unchanged with lapse of time, as a characteristic of a steady flow. [For correct indications, the extremity of the piezometer should have its edges flush with the inner face of the pipe wall, where it is inserted.] At $N_1$, although at the same level as $N_2$, we find, on inserting an open piezometer, $W_1$, that with $F_1 > F_2$ (and therefore with $\nu_1 > \nu_2$), $y_1$ is a little less than $y_2$; while if $F_1 < F_2$ (so that $\nu_1 < \nu_2$), then only is $y_1$ greater than $y_2$, but the difference is greater than before. We have therefore found experimentally that, in a general way, when water is flowing in a pipe it presses less against the side of the pipe than it did before the flow was permitted, or (what amounts to the same thing) the pressure between the transverse laminae is less than the hydrostatic pressure would be. In the portion $HN_2$ of the pipe we find the pressure less than one atmosphere, and consequently a manometer register, lagging pressures from zero upwards (and not simply the excess over atmospheric pressure, like the Bourdon steam gauge and the open piezometer just mentioned) must be employed. At $N_2$, e.g., we find the pressure $$= \frac{1}{2} \text{ atm}, \text{i.e., } \frac{P_3}{y} = 17 \text{ ft.}$$ Even below the level $BC$, by making the sections quite narrow (and consequently the velocities great) the pressure may be made less than one atmosphere. At the surface $BC$ the pressure is of course just one atmosphere, while that in the jet at $N_2$, entering the right-hand tank under water, is necessarily $p_2 = 1 \text{ atm.} + \text{press. due to col. } h'$ of water practically at rest; A diagram showing a pipe with a piezometer inserted into it. **STeady Flow of Liquids.** 651 i.e., $P = \frac{b}{h} + h$; (wheres if $N$, were stopped by a diaphragm, the pressure-head just on the right of the diaphragm would be $b + h$; and that on the left $b - A$.) Similarly, when a jet enters the atmosphere in parallel filaments its particles are under a pressure of one atmosphere, i.e., their pressure-head = $b = 34$ ft. (for water); for the air immediately around the jet may be considered as a pipe between which and the water is exerted a pressure of one atmosphere. **491. Recapitulation and Examples.—We have found experimentally, then, that in a steady flow of liquid through a rigid pipe there is at each section of the pipe a definite velocity and pressure which all the liquid particles assume on reaching that section; in other words, at each section of the pipe the liquid velocity and pressure remain constant with progress of time. EXAMPLE 1.—If in Fig. 582, the flow having become steady, the volume of water flowing in 3 minutes is found on measurement to be 194 cub. feet, the volume per second is, from eq. (1), § 490, $$Q = \frac{194}{3 \times 60} = 0.744 \text{ cub. ft. per second.}$$ EXAMPLE 2.—If the flow in 2 min. 20 sec. is 386.4 lbs., the volume of flow per second is [ft. lb. sec.; eqs. (1) and (2)] $$Q = \frac{V}{\sqrt{\frac{g}{t}}} = \frac{386.4}{\sqrt{\frac{32.5}{140}}} = 0.0441 \text{ cub. ft. per sec.}$$ EXAMPLE 3.—In Fig. 539 the height of the open piezometer at $N_1$ is $y_1 = 9$ feet; what is the internal fluid pressure? [Use the inch, lb., and sec.] The internal pressure is $$p_1 = p_a + g y_1 = 14.7 + 108 \times \frac{9}{32.5} = 18.6 \text{ lbs. per sq. inch.}$$ The pressure on the outside of the pipe is, of course, one atmosphere, so that the resultant bursting pressure at that point ($N_2$) is 3.9 lbs. per sq. in. EXAMPLE 4.—The volume of flow per second being .0441 652 MECHANICS OF ENGINEERING. cub. ft. per sec., as in Example 1, required the velocity at a section of the (circular) pipe where the diameter is 2 inches. [Use ft., lb., and sec.] $$v = \frac{Q}{F} = \frac{0.0441}{\frac{\pi(2)^2}{4}} = 2.02 \text{ ft. per sec.;}$$ while at another section of the pipe where the diameter is four inches (double the former) and the sectional area, $F$, is therefore four times as great, the velocity is $\frac{1}{4}$ of $2.02 = 0.505$ ft. per sec. 492. Bernoulli's Theorem for Steady Flow; without Friction. If the pipe is comparatively short, without sudden bends, elbows, or abrupt changes of cross-section, the effect of friction of the liquid particles against the sides of the pipe and against each other (as when eddies are produced, disturbing the parallelism of flow) is small, and will be neglected in the present analysis, whose chief object is to establish a formula for steady flow through a short pipe and through orifices. An assumption, now to be made, of flow in plane layers, or laminated flow, i.e., flow in lamina $\parallel$ to the axis of the pipe at every point, may be thus illustrated (Fig. 350), which shows steady flow proceeding through a pipe ($CD$ of indefinite extent). All the liquid particles which at any instant form a small lamina, or sheet, as $AB$, $\parallel$ to axis of pipe, keep company as a lamina throughout the whole flow. The thickness, $ds$, of this lamina remains constant as long as the pipe is of constant cross-section, but shortens up (as at $C$) as it passes into a smaller section, and lengthens out (as at $D$) in a part of the pipe where the section is narrower (i.e., the sectional area, $F$, is smaller). The mass of such a lamina is $Fds/v + g(Fds)$, its velocity at any section will be called $v$ (pertaining to that point of the pipe's axis), the pressure of the lamina just behind it is $Fp$, upon the rear face, while the resistance (at the same instant) offered by its neighbor just ahead is $F(p+dp)$ on the front face; also BERNOULLI'S THEOREM. 633 its weight is the vertical force $Fds^y$. Fig. 534 shows, as a free body, the lamina which at any instant is passing a point $A$ of the pipe axis. The velocity is $v$ and pressure $p$. Note well the forces acting: the pressures of the pipe wall on the edges of the lamina have no components in the direction of $v$, for the wall is considered smooth, i.e., those pressures are $\perp$ to wall; in other words, no friction is considered. To this free body apply eq. (7) of § 74, for any instant of any curvilinear motion of a material point $$\frac{dv}{dt} = \left( \text{tang. acceleration} \right) \times ds$$ (1) in which $ds$ = a small portion of the path, and is described in the time $dt$. Now the tang. acel. = $\mathscr{R}$ (tang. compn. of the acting forces) + mass of lamina, i.e, $$\text{tang. acc.} = \frac{Fp - F(v + dp)}{Fyd^y} + \frac{Fyd^y \cos \phi}{Fyd^y + g}$$ (2) Now, Fig. 535, at a definite instant of time, conceive the volume of water in the pipe to be subdivided into a great number of laminae of equal mass (which implies equal volumes A diagram showing a pipe with water flowing through it, divided into smaller sections. Fig. 535. in the case of a liquid, but not with gaseous fluids), and let the $ds$ just mentioned for any one lamina be the distance from its centre to that of the one next ahead; this mode of subdivision 654 MECHANICS OF ENGINEERING. make the $ds$ of any one lamina identical in value with its thickness $ds$, i.e., $$ds = ds \quad \ldots \quad \ldots \quad (3)$$ We have also $$ds \cos \phi = -ds, \text{ or } ds \cos \phi = -ds; \quad (4)$$ $s$ being the height of the centre of a lamina above any convenient horizontal datum plane. Substituting from (3), (3), and (4) in (1), we derive finally $$\frac{1}{g} w ds + \frac{1}{y} dp + ds = 0 \quad \ldots \quad (5)$$ The flow being steady, and the subdivision into laminae being of the nature just stated, each lamina in some small time $dt$ moves into the position which at the beginning of $dt$ was filled by the lamina next ahead, and acquires the same velocity, the same pressure on its faces, and the same value of $z$, that the front lamina had at the beginning of $dt$. Hence, considering the simultaneous advance made by all the laminae between two cross-sections $m$ and $n$, an equation like (5) for each of the laminae between any two cross-sections $m$ and $n$ of the pipe, thus obtaining an infinite number of equations, from which by adding corresponding terms, i.e., by integration, we obtain $$\frac{1}{g} \int_{m}^{n} v ds + \frac{1}{y} \int_{m}^{n} dp + \int_{m}^{n} dz = 0; \quad (6)$$ whence, performing the integrations and transposing, $$\frac{v_n^*}{2g} + \frac{P_n}{y} + z_n = \frac{v_m^*}{2g} + \frac{P_m}{y} + z_m \quad \text{\{ Bernoulli's Theorem\}} \quad (7)$$ Denoting by Potential Head the vertical height of any section of the pipe above a convenient datum level, we may state Bernoulli's Theorem as follows: In steady flow without friction, the sum of the velocity-head, pressure-head, and potential head at any section of the pipe is a constant quantity, being equal to the sum of the corresponding heads at any other section. APPLICATIONS OF BERNOULLI'S THEOREM. 658 It is noticeable that in eq. (7) each of the terms is a linear quantity, viz., a height, or head, either actual, such as $h_a$ or $h_b$, or ideal (all the others), but do not bring into account the absolute size of the pipe, nor even its relative dimensions ($v_a$ and $v_b$); however, are connected by the equation of continuity $F_a v_a = F_b v_b$, and contains no reference to the volume of water flowing per unit of time ($Q$) or the shape of the pipe's axis. When the pipe is of considerable length compared with its diameter the friction of the water on the sides of the pipe cannot be neglected (§ 512). It must be noticed that Bernoulli's Theorem does not hold unless the flow is steady i.e., unsteady conditions entering into the position just vacated by the one next ahead (of equal mass), comes also into the exact conditions of velocity and pressure in which the other was when in that position. [N.B. This theorem can also be proved by applying to all the water particles between $n$ and $m$, as a collection of small rigid bodies (water being incompressible) the theorem of Work and Energy for a collection of Rigid Bodies in § 142, eq. (xvi), taking the respective paths which they describe simultaneously in a single dt.] 458 Single Application of Bernoulli's Theorem without Friction. —Fig. 536 shows a large tank from which a vertical pipe of uniform section leads to another tank and dips below the surface of the water in the latter. Both surfaces are open to the air. The vessels and pipe being filled with water, and the lower end $m$ of the pipe unstopped, a steady flow is established almost immediately, the surface $BC$ being very large compared with $P$, the area of the (uniform) section of the pipe. Given $F$, $F'$, $A$, $A'$, heights $h_a$ and $h_b$, required the velocity $v_a$ of the jet at $m$ and also the pressure, $p_m$, at $a$ (in pipe near entrance of same). $m$ is in the jet, just clear of the pipe, and practically in the water-level, $AD$. The velocity $v_a$ is unknown, but the pressure $p_m$ is practically = $p_a$ = one atmosphere, since Diagram showing a large tank from which a vertical pipe leads to another tank and dips below the surface of water. Fig. 326 656 MECHANICS OF ENGINEERING. the pressure on the sides of the jet is necessarily the hydrostatic pressure due to a slight depth below the surface $AD$. :. Press.-Head at $m$ is $\frac{P_m}{y} = \frac{P_n}{y} = b = 34$ feet. . . . (§ 428) Now apply Bernoulli's Theorem to sections $m$ and $n$, taking a horizontal plane through $m$ as a datum plane for potential heads, so that $z_m = h$ and $z_n = 0$, and we have $$\frac{v_m^2}{2g} + b + 0 = \frac{v_n^2}{2g} + \frac{P_n}{y} + h$$ . . . . (1) But, assuming that the section of the pipe is filled at every point, we must have $$v_m = v_n;$$ for, in the equation of continuity $$F_{w,m}v_m = F_{w,n}v_n,$$ if we put $F_w = F_n$, the pipe being of uniform section, we obtain $v_m = v_n$. Hence eq. (1) reduces to $$\frac{P_n}{y} - b - h = 34 ft. - h$$ . . . . (2) Hence the pressure at $n$ is less than one atmosphere, and if a small tube communicating with an air-tight receiver full of air were screwed into a small hole at $n$, the air in the receiver would gradually be drawn off until the pressure in the tube was equal to $P_n$. This is the principle of Spengel's air-jump, mercury, however, being used instead of water, as for this heavy liquid $b =$ only 30 inches.) If $h$ is made > $b$ for water, i.e. > 34 feet (or > 30 inches for mercury), $p_n$ would be negative from eq. (2), which is impossible, showing that the assumption of full pipe-sections is not borne out. In this case, $h > b$, only a portion, $mv$, (in length somewhat less than $b$) of the tube will be kept full **APPLICATIONS OF BERNOUlli'S THEOREM.** 657 during the flow (Fig. 537); while in that part $K_n'$ vapor of water, of low tension corresponding to the temperature ($\S$ 489), will surround an internal jet which does not fill the pipe. As for the value of $v_a$, Bernoulli's Theorem, applied to $Rc$ and $m$, in Fig. 536, gives finally $v_a = \frac{2gh}{g}$. EXAMPLE. If $h$ is 20 feet, Fig. 536, and the liquid is water, the pressure-head at $a$ is (1t., lb., sec.) $$\frac{P_a}{y} = b - h = 34' - 20' = 14 \text{ ft},$$ and therefore $$p_a = 14 \times 62.5 = 875 \text{ lbs. per sq. ft.} = 6.07 \text{ lbs. per sq. in.}$$ 494. Second Application of Bernoulli's Theorem without Friction.—Knowing by actual measurement the open piezometer height $y_a$ at the section $n$ in Fig. 538 (so that the pressure-head, $\frac{P_a}{y} = b + y_a$, at $n$ is known); knowing also the vertical distance $h_a$ from $n$ to $m$, and the respective cross-sections $F_a$ and $F_m$ being the sectional area of the jet, flowing into the air, so that $\frac{P_a}{y} = b$, required the volume of flow per sec.; i.e., required $Q$, which $$= F_a v_a = F_m v_m \dots\dots (1)$$ The pipe is short, with smooth curves, if any, and friction will therefore be neglected. From Bernoulli's Theorem [eq. (7), $\S$ 492], taking $m$ as a datum plane for potential heads, we have $$\frac{v_m^2}{2g} + b + 0 = \frac{v_a^2}{2g} + (y_a + b) + h_m \dots\dots (8)$$ But from (1) we have 658 MECHANICS OF ENGINEERING. $$v_m = [F_m + F_s] v_m;$$ substituting which in (3) we obtain, solving for $v_m$, $$v_m = \frac{g(9g_0 + h_0)}{\sqrt{1 - \left(\frac{F_s}{F_m}\right)^2}}, \quad \ldots \ldots \quad (3)$$ and hence the volume per unit of time becomes known, viz. $$Q = F_m v_m, \quad \ldots \ldots \quad (4)$$ NOTE.—If the cross-section $F_s$ of the nozzle, or jet, is > $F_m$, $v_m$ becomes imaginary (unless $g_0$ is negative (i.e., $p_a < 0$ one atmos.), and numerically > $h_0$); in other words, the assigned cross-sections are not filled by the flow. EXAMPLE.—If $y_0 = 17$ ft. (thus showing the internal fluid pressure at $n$ to be $p_a = g(y_0 + b) = 14$ atmos., $b = 10$ ft., and that the jet is 4 inches in diameter at $m$ and 5 inches at the nozzle $m$, we have from (3) (using ft.-lb.-sec. system of units in which $\gamma = 32.2$) $$v_m = \frac{\sqrt{2 \times 32.2(17 + 10)}}{\sqrt{1 - \left(\frac{4^2}{4^2 + 1}\right)^2}} = 50.4\mathrm{~ft.~per~sec.}$$ [N.B. Since $F_s/F_m$ is a ratio and therefore an abstract number, the use of the inch in the ratio will give the same result as that of the foot.] Hence, from (4), $$Q = F_m v_m = 4\pi(F_s/F_m)^2 \times 50.4 = 2.474\mathrm{~cub.~ft.~per~sec.}$$ 495. Orifices in Thin Plate—Fig. 539. When efflux takes place through an orifice in a thin plate, i.e., a sharp-edged orifice in the plane wall of a tank, a contracted vein (or "vein" ORIFICE IN THIN PLATE. 659 contracta") is formed, the filaments of water not becoming parallel until reaching a plane, $m$, parallel to the plane of vessel wall, which for circular orifices is at a distance from the interior plane of vessel wall equal to the radius of the circular aperture. It is not clear whether this plane does the internal fluid pressure become equal to that of the surrounding medium (atmosphere, here), i.e., surrounding the jet. We assume unless the width of the orifice is small compared with $h$, unless the vessel wall is horizontal. The area of the cross-section of the jet at $n$, called the contracted section, is found on measurement to be from .80 to .84 of the area of the aperture with most orifices of ordinary shapes, even with widely different values of the area of aperture and of the height, or head, $h$, producing the flow. Calling this abstract number (.80 to .84) the Coefficient of Contraction, and denoting it by $C$, we may write $$F_m = CF_n$$ in which $F$ = area of the orifice, and $F_n$ = that of the contracted section. $C$ ranges from .60 to .64 with circular orifices, but may have lower values with some rectangular forms. (See table in § 503.) A lamina of particles of water is under atmospheric pressure at $n$ (the free surface of the water in tank or reservoir), while its velocity at $n$ is practically zero, i.e. $v_n = 0$ (the surface at $B$ being very large compared with the area of orifice). Its experiences increasing pressure as it slowly ceases until finally at $m$ it becomes a jet of uniform velocity when its velocity is rapidly accelerated and pressure decreased, in accordance with Bernoulli's Theorem, and its shape lengthened out, until finally at $m$ it forms a portion of a filament of a jet, its pressure is one atmosphere, and its velocity = $v_m$, we wish to determine. The course of this lamina we call a A diagram showing a cross-sectional view of a jet emerging from a thin plate. The diagram includes labels for various points and lines indicating the path of the jet and its relationship to the plate. Fig. 509. 659 660 MECHANICS OF ENGINEERING. "stream-line," and Bernoulli's Theorem is applicable to it, just as if it were enclosed in a frictionless pipe of the same form. Taking then a datum plane through the centre of $m$, we have $$\frac{p_m}{y} = b, \quad z_m = 0, \quad \text{and} \quad v_m = t;$$ while $$\frac{p_a}{y} \text{ also} = b, \quad z_a = h, \quad \text{and} \quad v_a = 0.$$ Hence Bernoulli's Theorem gives $$\frac{v_m^*}{2g} + b + 0 = 0 + b + h;$$ $$\therefore \frac{v_a^*}{2g} = h \dots \dots (1)$$ and $$v_m = \sqrt{2gh}.$$ That is, the velocity of the jet at $m$ is theoretically the same as that acquired by a body falling freely in vacuo through a height $= h$ - the "head of water." We should therefore expect that if the jet were directly vertically upward, as at $m$, Fig. 540, A diagram showing a vertical jet of water emerging from a pipe. The diagram includes labels for various points and measurements. would be actually attained. [See §§ 52 and 53.] Experiment shows that this is not so. The jet does not materially differ from $A$ if $h$ is not > 6 or 8 feet. For $h > 8$ ft., however, the actual height reached is < $h$, the difference being not only absolutely but relatively greater as $h$ is taken greater, since the resistance of the air is then more and more effective in depressing and breaking up the stream. (See § 578.) At $m$, Fig. 540, we have a jet under a head $= h'$, directed ORIFICE IN THIN PLATE. 661 at an angle $\alpha$, with the horizontal. Its form is a parabola (§ 81), and the theoretical height reached is $h' = A \sin \alpha$ (§ 80). The jet from an orifice in thin plate is very limpid and clear. From eq. (1), we have theoretically $$v_m = \sqrt{2gh}$$ (an equation we shall always use for efflux into the air through orifices and short pipes in the plane wall of a large tank whose water-surface is very large compared with the orifice, and is open to the air); but experiment shows that for an "orifice in thin plate" this value is reduced about 5% by friction at the edges, so that for ordinary practical purposes we may write $$v_m = v \sqrt{2gh} = 0.97 \sqrt{2gh}, \quad \ldots \quad (3)$$ in which $\phi$ is called the coefficient of velocity. Hence the volume of flow, $Q$, per time-unit will be $$Q = F_a v_m = CF\phi \sqrt{2gh}, \text{ on the average } = 0.02F \sqrt{2gh}. \quad (3)$$ It is to be understood that the flow is steady, and that the reservoir surface (very large) and the jet are both under atmospheric pressure. $\phi C$ is called the coefficient of efflux. EXAMPLE 1.—[Fig. 85.] Required the velocity of efflux, $v_m$, at an orifice in the plane wall of a tank, $Q_1$ per second, if $h = 21$ ft. 6 inches, the circular orifice being 2 in. in diam.; take $C = 0.64$. [Fl., lb., and sec.] From eq. (3), $$v_m = 0.97 \sqrt{2 \times 32.3 \times 21.5} = 36.1 \text{ ft. per sec.};$$ hence the discharge is $$Q = F_a v_m = 0.64 \times \frac{\pi}{12} \times 36.1 = 0.304 \text{ cub. ft. per second}.$$ EXAMPLE 2.—[Weibach.] Under a head of 3.396 metres the velocity $v_m$ in the contracted section is found by measure- 662 MECHANICS OF ENGINEERING. ments of the jet-curve to be 7.98 metres per sec., and the discharge proves to be 0.01835 cubic metres per sec. Required the coefficient of velocity ($\phi$) and that of contraction ($C$), if the area of the orifice is 36.3 sq. centimetres. Use the metre-kilogram-second system of units, in which $g = 9.81$ met. per sq. second. From eq. (2), $$\phi = \frac{v_m}{\sqrt{2gh}} = \frac{7.98}{\sqrt{2 \times 9.81 \times 3.396}} = 0.978;$$ while from (3) we have $$C = \frac{Q}{F_{ph} \cdot v_0 h} = \frac{Q}{F_{m} \cdot \sqrt{2gh} \times 7.98} = 0.631.$$ $\phi$ and $C$, being abstract numbers, are independent of the system of concrete units adopted. Note.—To find the velocity $v_m$ of the jet at the orifice by means of the jet-curve, as mentioned in Example 2, we may proceed as follows: Since we cannot very readily assure ourselves that the direction of the jet at the orifice is horizontal, we consider the angle $\alpha_o$ of the parabola (see Fig. 93 and § 80) as unknown, and therefore have two unknowns to deal with, and obtain the necessary two equations by measuring $x$ and $y$ (see page 84) of two points of the jet, remembering that if we use the equation (3) of page 84 in its present form, points of the jet below the orifice will have negative $y$. The first of these equations between $x$ and $y$, in equation (3) furnishes two equations between constants, in which only $\alpha_o$ and $h$ are unknown. To eliminate $\alpha_o$, for $$\frac{1}{\cos^2 \alpha_o},$$ we write $1 + \tan^2 \alpha_o$, and taking $x_o = 2h$, for convenience, we finally obtain $$A = \frac{1}{8} \left( x_o^2 + (4y - y_o)^2 \right),$$ and . . . $v_m = \sqrt{\frac{g(x_o^2 + (4y - y_o)^2)}{4(y_o - y)}}$$ in which $y$, and $y_o$ are the vertical distances of the two points. **BOUNDED ORIFICE.** 663 chosen below the orifice; that is, we have already made them negative in eq. (3) of page 84. The $A$ of the preceding equation simply denotes $v_{a} + 2g$, and must not be confused with that of the last two figures. For accuracy the second point should be as far from the orifice along the jet as possible. **408. Orifice with Rounded Approach.*—Fig. 541 shows the general form and proportions of an orifice or mouth-piece in the use of which contraction does not take place beyond the edges, the inner surface being one "of revolution," and so shaped that the liquid filaments are parallel on passing the outer edge $m$; hence $C = \frac{1}{2}$ (for water), $C = \frac{1}{3}$ (for mercury) (= 84 ft. for water and 30 inches for mercury) in Bernoulli's Theorem, if efflux takes place into the air. We have also the sectional area $F_m = F$ at that final edge of orifice, i.e., the coefficient of contraction, or $C_1 = unity = 1.00$, so that the discharge per time-unit has a volume $$Q = F_m v_m = F v_m$$ The tank being large, as in Fig. 540, Bernoulli's Theorem applied to $m$ and $n$ will give, as before, $$v_m = \sqrt{\frac{2gh}{\phi}}$$ as a theoretical result, while practically we write $$v_m = \phi \sqrt{\frac{2gh}{\phi}}, \quad \quad \quad \quad \quad \quad \quad (1)$$ and $$Q = F \phi \sqrt{\frac{2gh}{\phi}}. \quad \quad \quad \quad \quad \quad \quad (2)$$ As an average $\phi$ is found to differ little from 0.97 with this orifice, the same value as for an orifice in thin plate ($\S$ 405). **407. Problems in Efflux Solved by Applying Bernoulli's Theorem.—In the two preceding paragraphs the pressure-heads at sections $m$ and $n$ were each $= p_n + y =$ height of *Smooth conical nozzles for fire-steams give $\phi = .97$ with $h =$ pres.-head + veloc.-head at base of pipe; see p. 883.* A diagram showing a rounded approach orifice. 664 MECHANICS OF ENGINEERING. the liquid barometer = b; but in the following problems this will not be the case necessarily. However, efflux is to take place through a simple orifice in the side of a large reservoir, whose upper surface (n) is very large, so that $v_n$ may be put = zero. **Problem 1—Fig. 543.** What is the velocity of efflux, $v_m$, at the orifice m (i.e., at the contracted section, if it is an orifice in thin plate) of a jet of water from a steam-boiler, if the free surface at n is at a height = A above $v_m$, and the pressure of the steam at this point is $p_s$? The discharge taking place into the air. Applying Bernoulli's Theorem to section $m$ at the orifice [where the pressure-head is $b$ and velocity-head $v_m + \frac{2g}{y}$ (unknown)] and to section $n$ at water-surface [where velocity-head = 0 and pressure-head = $p_s + y$], we have, taking $m$ as a datum for potential heads so that $z_m = 0$ and $z_n = A$, $$\frac{v_m^2}{2g} + b + 0 = 0 + \frac{p_s}{y} + h.$$ $$\therefore v_m = \sqrt{\frac{2g(p_s - b + h)}{y}} \quad . . . . . . . . . . . . . . . . (1)$$ **EXAMPLE.—Let the steam-gauge read 40 lbs. (and hence $p_s = 54.7$ lbs. per sq. inch) and $h = 2$ ft. 4 in.; required $v_m$. Also if $F = \frac{1}{8}$ sq. in., in "thin plate," required the rate of discharge (volume). The temperature of saturated steam of the given tension must be 386° Fahr. [see foot of page 60]. The water is practically at the same temperature and hence of a heatiness, $\gamma$, of 57.7 lbs. per cubic ft. (p. 518). From eq. (1) above, then, with ft. lb. and sec., noting that for this case $b = [(14.7 + 2) \times 57.7] \text{ feet}$, $$v_m = \sqrt{\frac{2 \times 39.2 \times (54.7 \times 144 - 14.7 \times 144 - 95)}{57.7}}$$ $$= 81.1 \text{ ft. per sec., theoretically; but practically}$$ PROBLEMS OF EFFLUX—ORIFICES. 655 $$v_{m} = 0.97 \times 81.1 = 78.6 \text{ ft. per sec.};$$ so that the discharge begins at the rate of $$Q = 0.64 F v_{m} = 0.64 \times \frac{1}{3} \pi d^2 \times 78.6 = 0.174 \text{ cub. ft. p. sec.}$$ **Problem II—Fig. 548.** With what velocity, $v_m$, will water flow into the condenser $C$ of a steam-engine where the tension of the vapor is $p_m < 1$ atmosphere, if $h$ = the head of water, and the flow takes place through an orifice in thin plate? Taking position $m$ in the contracted section where the filaments are parallel, and the pressure is more equal than that of the sur- rounding vapor, we have, by application of Bernoulli's theorem (wide) free surface of the water at the tank, where (at surface) the pressure is one atmosphere [and : $\frac{F_m}{Y} = h = 84$ ft.] and velocity practically zero; we have, applying Bernoulli's Theorem to $n$ and $m$, taking $m$ as a datum level for potential heads (so that $s_n = A$ and $s_m = 0$), $$\frac{v_m^2}{2g} + \frac{F_m}{Y} + 0 = b + h,$$ $$v_m = \sqrt{2g(h + b - \frac{F_m}{Y})}, \quad \ldots \ldots (1)$$ and $$Q = F_m v_m, \quad \ldots \ldots (2)$$ as theoretical results. But practically we must write $$v_m = 0.97 \sqrt{2g(h + b - \frac{F_m}{Y})}, \quad \ldots \ldots (3)$$ and $$Q = F_m v_m = CFv_m, \quad \ldots \ldots (4)$$ in which $F =$ area of orifice in thin plate, and $C =$ coefficient of contraction = about 0.62 approximately [see § 493]. A diagram showing a steam engine with a condenser and an orifice in a thin plate. 068 MECHANICS OF ENGINEERING. **EXAMPLE.**—If in the condenser there is a "vacuum" of 37 inches (meaning that the tension of the vapor would support 37 inches of mercury, in a barometer), so that $$p_m = \left[\frac{1}{4} \times 14.7\right] \text{ lbs. per sq. inch, and } h = 12 \text{ feet},$$ while the orifice is $\frac{1}{4}$ inch in diameter; we have, using the ft., lb., and sec, $$v_m = 0.97 \sqrt{\frac{9 \times 32.9}{19 + 34 - \frac{1}{4} \times 14.7 \times 144}} = 51.1 \text{ ft. per sec.}$$ (We might also have written, for brevity, $$\frac{F_m}{y} = [34 : 30] \times 34 = 2.833,$$ since the pressure-head for one atmos. = 34 feet, for water. Hence, for a circular orifice in thin plate, we have the volume discharged per unit of time, $$Q = CF_0 = 0.62 \times \frac{\pi(\frac{1}{4})^2}{4 \times 12} \times 51.1 = 0.0431 \text{ cub. ft. per sec.}$$ **497a. Eflux through an Orifice in Terms of the Internal and External Pressures.—Fig. 544. Let efflux take place through a small orifice from the plane side of a large tank, in which st the level of the orifice the hydrostatic pressure was $p'$ before the opening of the orifice, that of the medium surrounding the jet being $p''$. When a steady flow is established, after opening the orifice, the pressure in the water on the level with the orifice will be materially changed only except in the immediate neighborhood of the orifice (see § 495); hence, applying Bernoulli's Theorem to $m$ in the jet, where the filaments are parallel, and a point $n$, in the body of the liquid and at the same level as $m$, and where the particles are practically at rest [i.e., $v_n = 0$] (hence not too near the) A diagram showing an orifice in a tank with water flowing out. Fig. 544 FORCE-PUMP. 667 orifice), we shall have, cancelling out the potential heads which are equal, $$\frac{v_{m}}{2g} + \frac{p'}{y} = \frac{0}{2g} + \frac{p'}{y},$$ ∴ $$v_m = 0.97 \sqrt{\frac{2g}{y}(\frac{p'}{y})}.$$ (1) (In Fig. 544 $p'$ would be equal to $p_1 + h_y$.) Eq. (1) is conveniently applied to the jet produced by a force-pump, supposing, for simplicity, the orifice to be in the head of the pump-cylinder, as shown in Fig. 545. Let the thrust (force) exerted along the piston-rod be $P$, and the area of the piston be $F'$. Then the intensity of internal pressure produced in the chamber $AB$ (when the piston moves uniformly) is $$p' = \frac{P + F'p_a}{F'},$$ while the external pressure in the air around the jet is simply $p_a$ (one atmos.) ∴ $$v_m = 0.97 \sqrt{\frac{2g}{y} \cdot \frac{P}{F'p_a}}.$$ (1') (N.B. Of course, at points near the orifice the internal pressure is $< p'$; read § 495.) **EXAMPLE.**—Let the force, or thrust, $P$, (due to steam-pressure on a piston not shown in figure,) be 2000 lbs., and the diameter of pump-cylinder be $d = 8$ inches, the liquid being salt water (so that $\gamma = 64$ lbs. per cubic foot). Then $$F' = \frac{4}{3}\pi(d^2) = 0.448 sq. ft.,$$ and [ft., lb., sec.] A diagram showing a force-pump with a piston moving through a cylinder containing a jet of water. Fig. 545. 688 MECHANICS OF ENGINEERING. $$u_m = 0.97 \sqrt{2 \times 32.2 \times \frac{9000}{0.448 \times 64}} = 65.4 \text{ ft. per sec}.$$ If the orifice is well rounded, with a diameter of one inch, the volume discharged per second is $$Q = F_a u_m = F_v v_m = \frac{\pi (1)^2}{4(12)} \times 65.4 = 0.353 \text{ cub. ft. per sec}.$$ To maintain steadily this rate of discharge, the piston must move at the rate [value = $v'$] of $$v' = Q + F'' = .353 + \left[\frac{\pi (9)^2}{4(12)}\right] = 0.800 \text{ ft. per sec},$$ and the force $P$ must exert a power ($\S$ 130) of $$L = Pw = 9000 \times 0.800 = 1600 \text{ ft. lbs. per sec}.$$ = about 3 horse-power (or 3 H. P.). If the water must be forced from the cylinder through a pipe or hose before passing out of a nozzle into the air, the velocity of efflux will be smaller, on account of "resistive friction" in the hose, for the same $F$, such a problem will be treated later ($\S$ 61). In this case, by the principle of analogy, by the use of several short-pistons, or of air-channels, practically steady flow is kept up, notwithstanding the fact that the motion of each piston is not uniform, and must be reversed at the end of each stroke. 498. Influence of Density on the Velocity of Efflux in the Last Problem.—From the equation $$u_m = \sqrt{2g\frac{p'' - p'}{y}}$$ of the preceding paragraph, where $p''$ is the external pressure around the jet, and $p'$ the internal pressure at the same level as the orifice but well back of it, where the liquid is ideally A diagram showing a cylinder with pistons and a nozzle. **RELATION OF DENSITY TO VELOCITY OF EFFLUX. 609** at rest, we notice that for the same difference of pressure $[p' - p']$ the velocity of efflux is inversely proportional to the square root of the heaviness of the liquid. Hence, for the same $(p' - p')$, mercury would flow out of the orifice with a velocity only $0.873$ of that of water; for $$\sqrt{\frac{62.5}{948}} = \sqrt{\frac{1}{13.5}} = \frac{272}{1000}$$ Again, assuming that the equation holds good for the flow of gases (as it does approximately when $p'$ does not greatly exceed $p''$; e.g., by 0 or 8 per cent), the velocity of efflux of atmospheric air, when at a heaviness of 0.873 lbs. per cub. foot, would be $$\sqrt{\frac{62.5}{0.087}} = \sqrt{775.3} = 27.8$$ times as great as for water, with the same $p' - p''$. (See § 548, etc.) **489. Efflux under Water. Simple Orifice—Fig. 546.** Let $h_+$ and $h_-$ be the depths of the (small) orifice below the levels of the "head" and "tail" waters respectively. Then, using the formula of § 427, we have for the pressure at a (at same level as $m$, the jet) $$p' = (h_+ + b)y,$$ and for the external pressure, around the jet at $m$, $$p'' = (h_- + b)y;$$ whence, theoretically, $$v_m = \sqrt{\frac{g'}{y}(p' - p'')} = \sqrt{2g'(h_+ - h_-)} = \sqrt{2gh}, \quad (1)$$ where $h =$ difference of level between the surfaces of the two bodies of water. A diagram showing a simple orifice in a water tank with a jet emerging from it. Fig. 546 670 MECHANICS OF ENGINEERING. Practically, $$v_m = \phi \sqrt{2gh}; \quad \ldots \ldots \quad (9)$$ out the value of $\phi$ for efflux under water is somewhat uncertain; as also that of $C$, the coefficient of contraction. Weisbach says that $\mu = \phi C$, is partly less for efflux into the air; others, that there is no difference (Trautwine). See also p. 389 of vol. 6, Jour. of Engin. Associations, where it is stated that with a circular mouth-piece of 0.37 in. diam., and of "nearly the form of the cessa contracta," $\mu$ was found to be .959 for discharge into the air, and .945 for submerged discharge. 500. Efflux from a Small Orifice in a Vessel in Motion. CASE I. When the motion is a vertical translation and uniformly accelerated.—Fig. 547. Suppose the vessel to move upward with a velocity $p$. (See § 8 above.) Taking $m$ and $a$ as in the two preceding paragraphs, we know that $p_m = p' - external\ pressure = one atom = p_a$ (and : $P_m = b$). As to the internal pressure at $n$ (same level as $m$, but well back of orifice), $p_n$, this is not equal to $(b + h)y$, because of the accelerated motion, but we can determine it by considering free the vertical column or prism of liquid between $n$ and $m$. The vertical forces acting on which are $p_dF_d$, downward at $O$, $p_dF_d$ upward at $n$, and its weight, $hdF_y$. All other pressures are horizontal. For a vertical upward acceleration = $\bar{p}$, the algebraic sum of the vertical components of all the forces must = mass × very. accel., i.e., $$dF(p_n - p_m - hy) = AydF/g \cdot \bar{p};$$ whence $$p_n = p_a + hy[1 + \frac{\bar{P}}{g}] \quad \ldots \ldots \quad (1)$$ Putting $p_a$ and $p_a'$ equal to the $p'$ and $p''$, respectively, of the equation, we have ORIFICE IN MOTION. 671 of § 497, $$v_m = \sqrt{2g(h - p_m)} \quad \ldots \ldots \ldots (3)$$ It must be remembered that $v_m$ is the velocity of the jet relative to the orifice, which is itself in motion with a variable velocity. The absolute velocity $v_a$ of the particles of the jet is found by the construction in § 83, being represented graphically half way between the ordinate of water level $w_m$, and the other velocity with which the orifice itself is moving at the instant, as part of the vessel. The jet may make any angle with the side of the vessel. On account of the flow the internal pressures of the water against the vessel are no longer balanced horizontally, and the latter will swing out of the vertical unless properly constrained. If $p = g = acc.$ of gravity, $v_a = \sqrt{2} \cdot \sqrt{2gh}$. If $p$ is negative and $= g$, $v_a = 0$; i.e., there is no flow, but both the vessel and its contents fall freely, without rotation. Case II. When both vessel and contents of the vessel both have a uniform rotary motion about a vertical axis with an angular velocity $\omega$ (§ 110). Orifice small, so that we may consider the liquid inside (except near the orifice) to be in relative equilibrium. Suppose the jet horizontal at $m$, Fig. 548. and the radial distance of the orifice from the axis to be $x$. The external pressure $p_a = p_m$, and the internal (see § 28), eqs. (5) and (6) is $$p_a = p_m + (h + x)\omega^2\frac{x}{2g}y$$ hence the velocity of the jet, relatively to the orifice, is, (from § 497, since $p_a$ and $p_m$ correspond to the $p'$ and $p''$ of that article), $$v_m = \sqrt{2g(\frac{p_a - p_m}{y})} = \sqrt{2gh + (\omega x)^2},$$ Fig. 548. **673** MECHANICS OF ENGINEERING. i.e., $$v_m = \sqrt{2gh + v^2} ; \quad \ldots \quad (3)$$ in which $w_e = ax$, = the (constant) linear velocity of the orifice in its circular path. The absolute velocity $v_a$ of the particles in the jet close to the orifice is the diagonal formed on $w$ and $v_a$ (§ 83). Hence by properly placing the orifice in the casing, $w_o$ may be made small or large, and thus the kinetic energy carried away in the effluent water be regulated, within certain limits. Equation (3) will be utilized subsequently in the theory of Barker's Mill. Example. If the blades make 100 revol. per min. (whence $\omega = 3\pi(100 - 60)$ radians per sec.), $h_1 = 12$ feet, and $x = 3$ ft.; then $(f_1 h_1 \text{ sec})$ $$v_m = \sqrt{9 \times 32.9 \times 12 + \left(\frac{2\pi(100 - 60)}{60}\right)^2} = 34.8 \text{ ft. per sec.}$$ (while, if the casing is not revolving, $v_m = \sqrt{2gh_1}$ = only 27.8 ft. per sec.). If the jet is now directed horizontally and backward, and also tangentially to the circular path of the centre of the orifice, its absolute velocity (i.e., relatively to the earth) is $$w_a = v_m - ax = 34.8 - 20.9 = 13.9 \text{ ft. per sec.},$$ and is also horizontal and backwards. If the volume of flow is $Q = 0.35$ cub. feet per sec., the kinetic energy carried away with the water per second (§ 138) is $$\frac{1}{2}Mw_a^2 = \frac{Qv}{g} \cdot \frac{w_a}{g} = \frac{1}{2} \times 62.5(13.9)^2 \times \frac{9}{32.9} = 46.8$$ ft. per second = 0.085 horse-power. **501. Theoretical Efflux through Rectangular Orifices of Considerable Vertical Depth, in a Vertical Plate.—If the orifice is so large vertically that the velocities of the different filaments in a vertical plane of the stream are theoretically different, having different "heads of water," we proceed as follows, taking into account, also, the velocity of approach, $v$, or mean velocity RECTANGULAR ORIFICES. 673 (If any appreciable), of the water in the channel approaching the orifice. Fig. 549 gives a section of the side of the tank and orifice. Let $a$ = width of the rectangle, the sill of the latter being horizontal, and $a = h_1 - h_2$, its height. Disregarding construction for the present, the theoretical volume of discharge per unit of time is equal to the sum of the volumes like $v_n dF$ (= $v_n b dz$), in which $v_n$ = velocity of any filament, as $m$, in the jet, and $b dz$ = cross-section of the small part which passes under any horizontal strip of the area of orifice, in a unit of time, its altitude being $v_n$. For each strip there is a different $x$ or "head of water," and hence a different velocity. Now the theoretical discharge (volume) per unit of time is $$Q = \text{sum of the volumes of the elem. prisms} = \int_{x=0}^{x=h} v_n dF;$$ i.e., $$Q = b \int_{x=0}^{h} v_n dx.$$ (1') But from Bernoulli's Theorem, if $k = c^2 + 2g$ is the velocity-head at $n$, the surface of the channel at head $C$, $b$ being the pressure-head at $n$, and $z$ its potential head referred to $m$ as datum (N.B. This $z = 84$ ft. for water, and must not be confused with the width $b$ of orifice), we have [see § 402, eq. (7)] $$\frac{v_n}{2g} + b + 0 = \frac{c^2}{2g} (\text{or } k) + b + z;$$ :. $v_n = \sqrt{2g} (\sqrt{x+k}) + b.$ (2) and since $dx = d(x-k)$, $k$ being a constant, we have, from (1') and (2), Theoret. $Q = b \sqrt{2g} \int_{h_1+k}^{h_1+k} \sqrt{x+k} dx.$ 674 MECHANICS OF ENGINEERING. or Theoret. $Q = \frac{4b}{3} \sqrt{\frac{2g}{h}} \left( (a_0 + k) - (a_0 + k') \right). \quad (1)$ (b now denotes the width of orifice.) If $c$ is small, the channel of approach being large, we have Theoret. $Q = \frac{4b}{3} \sqrt{\frac{2g}{h}} (a_0 - h_0) \quad \dots \quad (2)$ (c being $= Q +$ area of section of $nC$.) If $h_0 = 0$, i.e., if the orifice becomes a notch in the side, or an overfall [see Fig. 550, which shows the contraction which actually occurs in all these cases], we have for an overfall Theoret. $Q = \frac{4b}{3} \sqrt{\frac{2g}{h}} (a_0 - h_1) \quad \dots \quad (3)$ Note.—Both in (1) and (3) $h$, and $h_0$, are the vertical depths of the respective sills of the orifice from the surface of the water three or four feet back of the plane of the orifice, where the surface is comparatively level. This point must be specially attended to in deriving the actual discharge from the theoretical [see § 508]. If $Q$ were the unknown quantity in eqs. (1) and (3) it would be necessary to proceed by successive assumptions and approximations, since $Q$ is really involved in $k$; for $k = \frac{c'}{2g} \text{ and } F_c = Q$ (where $F_c$ is the sectional area of the channel of approach $nC$). With $k = 0$ (or $c$ very small, i.e., $F_c$ very large), eq. (3) reduces (for an overfall) to Theoret. $Q = \frac{4b}{3} \sqrt{\frac{2g}{h_1}} \quad \dots \quad (3')$ or $\frac{4b}{3}$ as much as if all parts of the orifice were at the same head of water = $h_1$ (as for instance if the orifices were in the horizontal bottom of a tank in which the water was $h_1$ deep, the orifices having a width = $b$ and length = $A_1$). Fig. 550 TRIANGULAR ORIFICE. 675 502. Theoretical Efflux through a Triangular Orifice in a Thin Vertical Plate or Wall. Base Horizontal.—Fig. 551. Let the channel of approach be so large that the velocity of approach may be neglected. $h_1$ and $h_2$ = depths of sill and vertex, which is downward. The analysis differs from that of the preceding article only in having $A = 0$ and the length $u$, of a horizontal strip of the orifice, variable; $b$ being the length of the base of the triangle. From similar triangles we have $$\frac{w}{b} = \frac{h_1 - x}{h_2 - h_1}, \text{i.e., } w = \frac{b}{h_2 - h_1}(h_2 - x).$$ ∴ Theoret. $Q = \int_{x=0}^{x=h_1} f(w) \cdot w dx = \int_{h_2-h_1}^{h_2} \frac{b}{h_2-h_1} (h_2-x) dx;$$ and finally, substituting from eq. (2) of § 501, with $k = 0$, Fig. 551. Fig. 552. Theoret. $Q = \frac{b}{h_2-h_1}\int_{x=0}^{x=h_1}(h_2-x)dx$ $$= \frac{9}{15} \frac{b^3}{h_2-h_1}\left[2(h_1-5h,h_1+3h,h)\right]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)$$ For a triangular notch as in Fig. 552, this reduces to Theoret. $Q = \frac{4}{15} h_1 \sqrt{\frac{3}{2}} b = \frac{8}{15} h_1 \sqrt{\frac{3}{2}} b; .$ . . . . . . . . . (5) i.e., $\frac{4}{15}$ of the volume that would be discharged per unit of 676 MECHANICS OF ENGINEERING. time if the triangular orifice with base $b$ and altitude $h_1$ were cut in the horizontal bottom of a tank under a head $h$, The measurements of $h_1$ and $b$ are made with reference to the level surface back of the orifice (see figure); for the water-surface in the plane of the orifice is below the level surface at the bottom of the tank. Prof. Thomson has found by experiment that with $b = 2h_1$, the actual discharge = theoret. disch. $\times 0.595$; and with $b = 4h_1$, actual = theoret. disch. $\times 0.680$. 503. Actual Discharge through Sharp-edged Rectangular Orifices (ells horizontal) in the vertical wall of a tank or reservoir. Case I. Complete and Perfect Contraction. The actual volume of water discharged per unit of time is much less than the theoretical values derived in § 501, chiefly on account of contraction. By complete contraction we mean that no edge of the orifice is flush with the side or bottom of the reservoir; and by perfect contraction, that the amount of spillage, when the heads $h_1$ and $h_2$ are measured, is so large that the contraction is practically the same if the channel were of infinite extent sideways and downward from the orifice. For this case ($a$, not zero) it is found most convenient to use the following practical formula ($b$ = width): $$Q = \mu_{ab} \sqrt{2g\left[h_1 + \frac{a}{2}\right]}$$ (6) in which (see Fig. 553) $a$ = the height of orifice, $h_1$ = the vertical depth of the upper edge of the orifice below the level of the reservoir surface, measured a few feet back of the plane of the orifice, and $\mu_a$ is a coefficient of orifice (an abstract number), dependent on experiment. With $a$, very approximate results (within 3 or 4 per cent) may be obtained from eq. (6) with openings not more than Fig. 553 503 RECTANGULAR ORIFICES. 677 18 inches, or less than 1 inch, high; and not less than 1 inch wide; with heads $\left(\frac{a}{2} + \frac{a}{2}\right)$ from 1 ft. to 20 or 30 feet. **Example.**—What is the actual discharge (volume) per minute through the orifice in Fig. 553, 14 inches wide and 1 foot high, the upper sill being 8 ft. 6 in. below the surface of still water? Use eq. (6) with the ft., lb., and sec. as units, and $\mu_4 = 0.02$. Solution: $$Q = 0.62 \times 1 \times 14 \times \sqrt{92} \times 84 + 4 = 17.41 \text{ cubic feet per second},$$ while the flow of weight is $$G = Qv = 17.41 \times 62.5 = 1088 \text{ lbs. per second}.$$ Poncelet and Leibov's Experiments.—For comparatively accurate results, values of $\mu_4$ taken from the following table (computed from the careful experiments of Poncelet and Leibov), may be used for all orifices given in this book, provided practicable, for other sizes by interpolation. To use the table, the values of $h$, $a$, and $b$ must be reduced to metres, which can be done by the reduction-table below; but in substituting in eq. (6), if the metre-kilogram-second system of units be used $g$ must be put $=9.81$ metres per square second (see § 51), and $Q$ will be obtained in cubic metres per second. Since $\mu_4$ is an abstract number, once obtained as indicated above, it does not necessitate any particular system of units in making substitutions in eq. (6). The ft., lb., and sec will be used in all cases. ',
1 foot = 0.30479 metres. 1 inch = 0.0253 metres.
3 feet = 0.09095 2 inches = 0.0507
5 feet = 0.15195 3 inches = 0.0691
4 " = 1.3918 metres. 4 " = 0.1015
5 " = 1.53937 5 " = 0.1968
6 " = 1.58783 6 " = 0.2789
7 " = 1.63665 7 " = 0.3776
8 " = 2.43836 8 " = 0.5030
9 " = 2.74315 9 " = 0.58283
10 " = 3.03794 10 " = 0.67994
TABLE FOR REDUCING FEET AND INCHES TO METRES.
"'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "'', "', " "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n", "'\n, " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " " "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" 678 MECHANICS OF ENGINEERING. TABLE, FROM PONCELET AND LESBOS. VALUES OF $\mu_0$ FOR Eq (6), FOR RECTANGULAR ORIFICES IN THIS PLATE. (Complete and perfect contraction.)
Value of $a$, ft. $b = 0.00$ in. $b = 0.05$ in. $b = 0.10$ in. $b = 0.20$ in. $b = 0.30$ in. $b = 0.40$ in. $b = 0.50$ in. $b = 0.60$ in. $b = 0.70$ in. $b = 0.80$ in.
0.015 P1 P2 P3 P4 P5 P6 P7 P8 P9 P10
0.019 0.607 0.636 0.665 0.694 0.723 0.752 0.781 0.810 0.839 0.868
0.157 0.993 1.012 1.031 1.051 1.071 1.091 1.111 1.131 1.151 1.171
0.388 2.249 2.278 2.307 2.336 2.365 2.394 2.423 2.452 2.481 2.510
0.595 3.585 3.614 3.643 3.672 3.701 3.730 3.759 3.788 3.817 3.846
0.799 4.999 5.028 5.057 5.086 5.115 5.144 5.173
                                                                     <table cellspacing="none" cellpadding="none"> <tr> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> <th>P</th> </tr> <tr> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> <th>$\mu_0$</th> </tr> <tr style="background-color: #f2f2f2;"> <th colspan="11">$\mu_0 = 4 \text{ in}$.</table cellspacing="none" cellpadding="none">
> >
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
>
679 From the foregoing table, for $a = 0.10^{\circ}$, $b = 0.80^{\circ}$ and $a = 0.30^{\circ}$, we find $\mu_a = 0.692$ " $A = 0.10^{\circ}$, $b = 0.80^{\circ}$ " $a = 0.20^{\circ}$, " $\mu_a = 592$ diff. = .010 Hence, by interpolation, for $A = 0.10^{\circ}$ $b = 0.51^{\circ}$, and $a = 0.20^{\circ}$, we have $\mu_a = 0.609 - \frac{1}{\sqrt{2}} [0.809 - 0.592] = 0.600$ Hence [ft., lb., sec.] remembering that $\mu_a$ is an abstract number, from eq. (6), $$Q = 0.600 \times \frac{1}{\sqrt{2}} \times \frac{1}{\sqrt{2}} \times 32.2 (\frac{A}{\pi} + \frac{b}{\pi}) = 4.36$$ cub. ft. per second. **Case II. Incomplete Contraction.—This name is given to the cases, like those shown in Fig. 554, where one or more sides of the orifice have an interior border flush with the sides or bottom of the (square-cornered) tank. Not only is the general direction of the stream altered, but the discharge is greater, on account of the larger size of the contracted section, since contraction is prevented on those sides which have a border. It is assumed that the contraction which does occur (on the other edges) is perfect; i.e., the cross-section of the water is harmonious with the orifice. According to the experiments of Bidond and Weibach with Poncelet's orifices (i.e., orifices in thin plate mentioned in the preceding table), the actual volume discharged per unit of time is $$Q = \mu ab \sqrt{2g(b + \frac{a}{2})} \cdot (7)$$ (differing from eq. (6) only in the coefficient of efflux $\mu$, in which the abstract number $\mu$ is found thus: Determine a coefficient of efflux $\mu_a$ as if eq. (6) were to be used in Case I; i.e., as if contraction were complete and perfect; then write) A diagram showing a rectangular orifice with a border flush with the sides. Fig. 554 680 MECHANICS OF ENGINEERING. $$\mu = \mu_0 [1 + 0.155 n], \quad \ldots \quad (7)$$ where $n$ is the ratio of the length of periphery of the orifice with a border to the whole periphery. E.g., if the lower sill, only, has a border, $$n = b - [2(a + b)]$$ while if the lower sill and both sides have a border, $$n = (2a + b) + [2(a + b)].$$ **Example.**—If $h_1 = 8$ ft. ($= 2.43m$), $b = 2$ ft. ($= 0.60m$), $a = 4$ in. ($= 0.10m$), and one side is even with the side of the tank, and the lower sill even with the bottom, required the volume discharged per second. (Sharp-edged orifice, in vertical plane, etc.) Here for complete and perfect contraction we have, from Poole's tables (Case I), $\mu_0 = 0.608$. Now $n = \frac{1}{4}$; hence, from eq. (7), $$\mu = 0.608 [1 + 0.155 \times \frac{1}{4}] = 0.6551;$$ hence, eq. (7), $$Q = 0.655 \times 2 \times \frac{1}{4} \times 92.2(8 + 1) = 10.28 \text{ cwt. ft. per sec}.$$ **Case III. Imperfect Contraction.**—If there is a submerged channel of approach, symmetrically placed as regards the orifice, and of an area (cross-section) $G$, not much larger than that, $F$, of the orifice (see Fig. 55), the contraction is less than in Case I, and is called imperfect contraction. Upon his experiments with Poole's orifices, with imperfect contraction, Weidbach bases the following formula for the discharge (volume) per unit of time, viz, $$Q = \mu_{ab} \sqrt{9p(h + \frac{a}{2})}, \quad \ldots \quad (8)$$ Fig. 55. RECTANGULAR ORIFICES. 681 (see Fig. 553 for notation), with the understanding that the coefficient $$\mu = \mu_c (1 + \beta) \quad \quad \quad (8')$$ where $\mu_c$ is the coefficient obtained from the table of Case I (as if the contraction were perfect and complete), and $\beta$ an abstract number depending on the ratio $F : Q = m$, as follows: $$\beta = 0.0760 [9'' - 1.00] \quad \quad \quad (8'')$$ To shorten computation Weisbach gives the following table for $\beta$: **EXAMPLE.**—Let $h_1 = 4^{\prime\prime} \times 4^{\prime\prime} = 1.46$ met.), the dimensions of the orifice being—
$m$ $\beta$ $m$ $\beta$
width = $b = 8$ in. (= 0.90"); .05 .009 .85 .178
height = $a = 5$ in. (= 0.128"); .15 .080 .85 .341
while the channel of approach ($CD$, Fig. 555) is one foot square. From Case I, we have, for the given dimensions and head, .30 .030 .85 .278
.30 .071 .85 .865
.40 .107 .85 .438
.50 .158 1.00 .608
$\mu_c = 0.610$;
$F/Q = \frac{40}{144}$ sq. in. = 0.27.
We find [Table A] $$\beta = 0.062;$$ and hence $\mu = \mu_c (1.062)$, from eq (8'). Therefore, from eq. (8), with ft., lb., and sec, $$Q = 0.610 \times 1.062 \times \sqrt{F} \times \sqrt{Q} \times 32.9 \times 5$$ = 3.92 cub. ft. per sec. **Case IV. Head measured in Moving Water.—See Fig. 556.** If the head $A_1$, of the upper sill, cannot be measured to the level of still water, but must be taken to the surface of a channel of approach, where the velocity of approach is quite 682 MECHANICS OF ENGINEERING. appreciable, not only is the contraction imperfect, but strictly we should use eq. (1) of § 501, in which the velocity of approach is considered. Let $F = \text{area of orifice},$ and $G = \text{diameter of orifice}.$ When the orifice is closed, then the velocity of approach is $c = Q/G,$ and $k$ (of above eq.) $= c^2 + 2gQ = Q^2 + 2gC^2;$ but $Q$ itself being unknown, a substitution of $k$ in terms of $Q$ in eq. (1), § 501, leads to an equation of high degree with respect to $Q.$ Practically, therefore, it is better to write $$Q = \mu\sqrt{2g(h_1 + \frac{a}{2})} \quad \ldots \quad (9)$$ and determine $\mu$ by experiment for different values of the ratio $F/G.$ Accordingly, Weibach found, for Poncelet's orifices, that if $\mu_c$ is the coefficient for complete and perfect contraction from Case I, we have $$\mu = \mu_c(1 + \beta). \quad \ldots \quad (9')$$ $\beta'$ being an abstract number, and being thus related to $F/G,$ $$\beta' = 0.641(\frac{F}{G}). \quad \ldots \quad (9'')$$ A was measured to the surface one meter back of the plane of the orifice, and $F'/G$ did not exceed 0.50. Weibach gives the following table computed from eq. (9''):
$F/G$$\beta'$
0.050.008
0.100.008
0.150.008
0.200.008
0.250.008
0.300.008
0.350.008
0.400.079
0.450.139
0.500.186
Example.--A rectangular water-trough 4 ft wide is dammed up with a vertical board which is a rectangular orifice, as in Fig. 556, of width $b = 2$ ft. (= 6.06 met.), and height $a = e$ in. (= 0.15 met.), and when the water-level behind the board has ceased rising (i.e., when the water has risen some $\delta(e/a)$), we find that $h_1 = 2$ ft., and the depth behind the trough to be 4 ft. Required $Q.$ Since $F' = 1$ sq. ft., $\beta' = f_1$, we find (Table B) $\beta' = 0.005;$ and $\mu_c$ being = 0.182 from Poncelet's tables, Case I, we have finally, from eq. (9), $$Q = \mu\sqrt{2g(h_1 + \frac{a}{2})}. \quad \ldots \quad (9)$$ DISCHARGE OF OVERFALL-WEIRS. 683 $$Q = 0.619(1.005) \times \frac{1}{2} \times 32.5 \times 2.55 = 7.41 \text{ cubic feet per second.}$$ **504. Actual Discharge of Sharp-edged Overfalls (Overfall- weirs; or Rectangular Notches in a Thin Vertical Plate).** Case I. Complete and Perfect Contraction (the normal case), Fig. 557; i.e., no edge is flush with the side or bottom of the reservoir, whose sectional area is very large compared with that, $A_1$, of the notch. By depth, $h_1$, of the notch, we can understand the depth of water will be below the level of a few feet back of the notch where it is level. In the plane of the notch the vertical thickness of the stream is only from $\frac{1}{4}$ to $\frac{1}{8}$ of $A_1$. Putting, therefore, the velocity of approach = zero, and hence $k = 0$, in eq. (5) of § 501, we have for the actual discharge: $$Q = \mu_1 A_1 h_1 \sqrt{2g} h_1 \ldots \ldots \ldots (10)$$ ($b$ = width of notch), where $\mu_1$ is a coefficient of efflux to be determined by experiment. Experiments with overfalls do not agree as well as might be desired. Those of Poncelet and Leibnitz gave the results in Table C. **EXAMPLE 1.—With**
For $b = 0.00\%$ For $b = 0.05\%$
inches. inches. inches. inches.
$h_1$ = 1 ft. 4 in. (= .400\") 01 .696 .638 .638
$b$ = 2 ft. (= .60\") 02 .818 .799 .799
We have, from Table C, $\mu_1 = .586$, and $(ft., lb., sec.)$ 03 .818 .809 .809
$\therefore Q = .586 \times \frac{1}{2} \times 32.5 \times \sqrt{2g} h_1$ 04 .809 .809 .809
= 9.54 cubic feet per second. 05 .809 .809 .809
06 .809 .809 .809
07.809.809.809
08.809.809.809
09.809.809.809
10.809.809.809
11.809.809.809
12.809.809.809
13.809.809.809
14.809.809.809
15



























































684 MECHANICS OF ENGINEERING. Since $b$ is unknown, we cannot use the table immediately, but take $\mu_c = .600$ for a first approximation; whence, eq. (10), (ft., lb., sec.) $$b = \frac{6}{0.6 \times 4 \times 12 \times 32.2 \times \frac{1}{12}} = 2.46 \text{ ft}.$$ Then, since this width does not much exceed 0.60 metre, we may take, in Table C, for $A_b$, for $\mu_c = .589$; $$b = \frac{6}{.589 \times 4 \times 12 \times 32.2 \times \frac{1}{12}} = 2.50 \text{ ft}.$$ **Case II. Incomplete Contraction; i.e., both ends are flush with the sides of the tank, those being $b$ to the plane of the notch. According to Weishab, we may write** $$Q = \frac{4\pi b h_1 + 2gh_1}{g}h_2,$$ (11) in which $\mu = 1.041\mu_c$, $\mu_c$ being obtained from Table C for the normal case, i.e., Case I. The section of channel of approach is large compared with that of the notch; if not, see Case IV. **Case III. Imperfect Contraction; i.e., the velocity of approach is appreciable; the sectional area $G$ of the channel of approach not being much larger than that, $F = bh_1 = area$ of notch. Fig. 558. $b$ is width, and $h_1$ is depth of notch (see Case I). Here, instead of using a formula involving** $$k = c^* + 2g = [Q + G]^* + 2g$$ (see eq. (3), § 501), it is more convenient to put $$Q = \frac{4\pi b h_1 + 2gh_1}{g}h_2,$$ as before, with $$\mu = \mu_c(1 + \beta)$$ (18) in which $\mu_c$ is for the normal case [Case I]; and $\beta$, according OVERFALL-WEIBS. 685 to Weibach's experiments, may be obtained from the empirical formula $$\beta = 1.718 \left( \frac{F}{Q} \right)$$ (12) [Table D is computed from (12).] (The contraction is complete in this case; i.e., the ends are not flush with the sides of the tank.) EXAMPLE.—If the water in the channel of approach has a vertical transverse section of $A = 9$ sq. feet, with both ends flush with the sides (i.e., $b = 9$) and 1 foot deep ($h = 1$) (to level of surface of water 3 ft. or less of bottom), we have, from Table C, with $b = .60$ met. and $h_a = 0.30$ met. $$\mu = 0.586$$ while from Table D, with $F$: $Q = 0.999$ (or $\frac{4}{3}$), $$\beta = .005$$ hence (ft.-lb.-sec. system of units), from eq. (12), $$Q = 1 \times 0.586 \times 1.005 \times 2 \times 1 \times \sqrt{64.4 \times 1.0}$$ = 6.30 cub. ft. per second. CASE IV. Fig. 559. Imperfect and incomplete contractions together; both end-constructions being "suppressed" (by making the ends flush with the sides of the reservoir, these sides being at a distance of about one foot from the mouth), and the channel of approach not being very deep, i.e., having a sectional area $Q$ but little larger than that, $F$, of notch: $F = b h_a$ as before. Again we write $$Q = \frac{4}{3} \mu b h_a \sqrt{2gh_a}$$ with $\mu$ computed from $$\mu = \mu_1 (1 + \beta)$$ (13) $\mu$ being obtained from Table C; while 686 MECHANICS OF ENGINEERING. $$\beta = 0.041 + 0.3693 \left( \frac{F}{G} \right)^* \dots \dots (13^*)$$ an empirical formula based by Weibach on his own experiments.* To save computation, $\beta$ may be found from Table E, founded on eq. (13)*.
$F$ 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
$G$ 0.041 0.043 0.045 0.049 0.056 0.064 0.074 0.086 0.100 0.116 0.133
**EXAMPLE.—Fig. 559. With** $$b = 2 \text{ ft. (= 0.60 met.)}$$ and $$h_s = 1 \text{ ft. (= 0.30 met.)}$$ we have, from Table C, $\mu_s = 0.556$. But, the ends being flush with the sides of the reservoir or channel, and $G$ being $= 6$ sq. ft. (see figure), which is not excessively large compared with $F = b h_s = 2$ sq. ft., we have, from Table E, with $F$: $G = 0.333$, $$\beta = .081;$$ and hence [eq. (13) and (13')] $\mu_s$ being $.586$ as in last example, $$Q = \frac{4 \times 0.586 \times (1 + .891) \times 2 \times 1 \times .644 \times .16}{6.78} \text{ cfs./ft.\ per sec.}$$ 585. Francis' Formula for Overfall (i.e., rectangular notch). —from extensive experiments at Lowell, Mass., in 1851, with rectangular overfall-weirs, Mr. J. B. Francis deduced the following formula for the volume, $Q$, of flow per second over such weirs 10 feet in width, and with $h_s$ varying from $6$ to $1.6$ feet (from sill of notch to level surface of water a few feet back): * Weibach's results in this case differ considerably from those of Basin; see p. 688. Fig. 559. OVERFALL-WEIRS. 687 $$Q = \frac{4}{3} \times 0.6224 (\beta - \frac{1}{\sqrt{n}}h) \cdot \frac{2gL}{b}, \quad . . . (14)$$ In which $b$ = width. This provides for incomplete contraction, as well as for complete and perfect contraction, by making $n = 2$ for perfect and complete contraction (Fig. 557); $n = 1$ when one end only is flush with side of channel; $n = 0$ when both ends are flush with sides of channel. The contraction was considered complete and perfect when the channel of approach was made as wide as practicable, $= 13.96$ feet, the depth being about 5 feet. Mr. Francis also experimented with submerged or "drowned" weirs in 1883; such a weir being one in which the sill is below the level of the tail-water (i.e., of receiving channel). 508. Falesley and Stearn's Experiments at Boston, Mass., in 1877 and 1880.—These may be found in the Transactions of the American Society of Civil Engineers, vol. xxi, and gave rise to formulae differing slightly from those of Mr. Francis in some particulars. In the case of supposed end-contractions, like that in Fig. 359, they propose formulae as follows: When depth of notch is not large, $$Q \text{(in cub. ft. per sec.)} = 3.31 b_1 h + 0.007 b \quad . . . (15)$$ ($b$ and $h$, both in feet), "A $_4$, the depth on the weir, should be measured from the surface of the water above the curvature of the sheet." "Air should have free access to all space under the sheet." The cross-section of the sheet is shown in Fig. 360. The formula does not apply to depths on the weir less than 0.07 feet. When the depth of notch is quite large, a correction must be made for velocity of approach, $c$, thus: $$Q \text{(in cub. ft. per sec.)} = 3.31 b \left[ A_4 + 1.5 \frac{c^2}{g} \right] + 0.007 b \quad . . . (16)$$ ($b$ and $h$, both in feet). 688 MECHANICS OF ENGINEERING. The channel should be of uniform rectangular section for about 30 ft. more from the weir, to make this correction properly. If $G$ = the cross-section, in sq. ft., of the channel of approach, $Q$ is found approximately by dividing an approximate value of $Q$ by $G$; and so on for closer results. The weir may be of any length, $L$, from 5 to 19 feet. 506a. Recent Experiments on Overfall-weirs in France.—In the Annales des Ponts et Chaussées for October 1888 is an account of extensive and careful experiments conducted in 1886 and 1887 by M. Bazin on the flow over shallow overfall-weirs with end-conditions similar to those like that shown in Fig. 529. The widths of the weirs ranged from 0.50 to 2.00 metres, and the depths on the weirs ($h_1$) from 0.05 to 0.60 metre. With $p$ indicating the height of the sill of the weir from the bottom of the channel of approach, M. Bazin, as a practical result of the experiments, recommends the following formula as giving a reasonably accurate value for the volume of discharge per unit of time: $$Q = \frac{4}{\mu} \left[1 + \frac{0.55(h_1 - p)}{p + h_1}\right]^{2} \left( h_1 + \frac{0.0148}{p} \right) \dots (17)$$ where the coefficient $\mu'$ has a value $$\mu' = 0.6075 + \frac{0.0148}{h_1 (\text{in ft.)}} \dots (18)$$ Eq. (17) is homogeneous, i.e., admits of any system of units. Provision was made in these experiments for the free entrance of air under the sheet (a point of great importance), while the pressures at all points upstream were continued down-stream, beyond the plane of the weir, to allow any lateral expansion of the sheet. The value of $p$ ranged from 0.20 to 2.00 metres. Herr Ritter von Wex in his "Hydrodynamik" (Leipsic, 1886) derives formulae for weirs, in the establishing of which some rather peculiar views in the Mechanics of Fluids are advanced. OVERFALLS. 689 Formulae and tables for discharge through orifices or over weirs of some forms not given here may be found in the works of Weibach, Rankine, and Troutwine. Mr. Hamilton Smith, a noted American hydraulic engineer, has recently published "Hydraulics," a valuable compilation and record of the author's many experiments in all fields of hydraulics (New York, 1886: John Wiley & Sons). 507. Eflux through Short Cylindrical Tubes.—When efflux takes place through a short cylindrical tube, or "short pipe," at least 24 times as long as wide, inserted at right angles in the plane side of a large reservoir, the inner corners are rounded (see Fig. 386), so that jets issue from the tube in parallel filaments and with a sectional area, $F_m$, equal to that, $F_t$ of interior of tube. To attain this result, however, the tube must be full of water before the outer end is unstopped, and must not be oily; nor must the head, $h$, be greater than about 40 ft. for efflux into the air. Since at $m$ the filaments are parallel and the pressure head therefore equal to $(c = 34 \text{ ft. of water})$, the value of $\phi_0$ is zero; i.e., head due to one atmosphere in this instance; an application of Bernoulli's Theorem [eq. (7), § 499] to positions $m$ and $n$ would give (precisely as in §§ 454 and 455) $$v_m = \text{veloc. at } m = \sqrt{2gh}$$ as a theoretical result; but experiment shows that the actual value of $v_m$ in this case is $$v_m = \phi_0 \cdot \sqrt{2gh} = 0.815 \cdot \sqrt{2gh} \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)$$ 0.815 being an average value for $\phi_0$, the coefficient of velocity, for ordinary purposes. It increases slightly as the head decreases, 260 MECHANICS OF ENGINEERING. and is evidently much less than the value 0.97 for an orifice in a thin plate, § 495, or for a rounded mouth-piece as in § 496. But as the sectional area of the stream where the filaments are parallel, at $m$, where $v_m = 0.815 \frac{F}{2gh}$, is also equal to that, $F'$ of the tube, the coefficient of efflux $\mu_s$, in the formula $$Q = \mu_s F' \sqrt{2gh}$$ is equal to $\phi_s$; i.e., there is no contraction, or the coefficient of contraction, $C_s$ in this case is 1.00. Hence, for the volume of discharge per unit of time, we have practically $$Q = \phi_s F' \sqrt{2gh} = 0.815 F' \sqrt{2gh}.$$ (9) The discharge is therefore about $\frac{3}{4}$ greater than through an orifice of the same diameter in a thin plate under the same head [compare eq. (8), § 495]; for although at $m$ the velocity is less in the present case, the sectional area of the stream is greater, there being no contraction. This difference arises due principally to the fact that the entrance of the tube has square edges, so that the stream contracts (at $m'$, Fig. 561) to a section smaller than that of the tube, and then re-expands to the full section, $F'$ of tube. The eddying and accompanying internal friction caused by this re-expansion (or "sudden enlargement" of the stream) is the prin- cipal resistance which diminishes the velocity. It is noticeable also, in this case that the jet is not limpid and clear, as from thin plate, but troubled and only transluent (like ground-glass). The internal pressure in the stream at $m$ is found to be less than one atmosphere, viz., about 0.8 atm., as shown experimentally by the sudden ing in of air when a small aperture is made in the tube op- Fig. 561. Fig. 562. INCLINED SHORT PIPES. 091 poste $m'$. If the tube itself were so formed internally as to fit this contracted vent, as in Fig. 563, the eddy would be diminished and the velocity at $m$ increased, and hence the volume $Q$ of efflux increased in the same proportion. (See § 509.) If the tube is less than 24 times as long as wide, or if the interior is not wet by the water (as when greasy), or if the head is over 40 or 50 ft. (about), the efflux takes place as if the tube were not there, Fig. 563, and we have $$v_m = 0.97 \sqrt{gh}, \text{ as in } \S 495.$$ **EXAMPLE.** The discharge through a short pipe 3 inches in diameter, like that in Fig. 560, is 30 cubic feet per minute under a head of $g^{\prime}h$, say 10 feet. Required the coefficient of efflux $\mu_e = \phi_e$, in this case. For variety use the inch-pound-minute system of units, in which $g = 32\frac{1}{2} \times 12 \times 3600$ (see Note, § 51). $\mu_e$, being an abstract number, will be the same numerically in any system of units. From eq. (2), $$\phi_e = \mu_e = \frac{Q}{F^{\prime}\sqrt{gh}} = \frac{90 \times 1798}{\pi \times 3^2 \times g^{\prime} \times 32\frac{1}{2} \times 12 \times 60^{\circ} \times 30}$$ $$= 0.803.$$ § 508. Inclined Short Tubes (Cylindrical).—Fig. 564. If the short tube is inclined at some angle $\alpha < 90^{\circ}$ to the interior plane of the reservoir wall, the efflux is smaller than when the angle is $90^{\circ}$, as in § 507. We still use the form of equation $$Q = \mu F^{\prime}\sqrt{gh} = \phi F^{\prime}\sqrt{gh}, \quad (3)$$ but from Weibach's experiments $\mu$ should be taken from the following table: A diagram showing an inclined short tube with a vent near its base. Fig. 564 692 MECHANICS OF ENGINEERING. TABLE F. COEFFICIENT OF EFFLUX (INCLINED TUBE).
For α = 90° 80° 70° 60° 50° 40° 30°
Take µ = φ = .715 .799 .769 .764 .747 .731 .719
EXAMPLE.--With h = 12 ft., d = diam. of tube = 4 ins., and α = 45°, we have for the volume discharged per sec. (ft., lb., and sec.) $$Q = [0.731 + \frac{1}{4}(.016)\frac{\pi(1)}{3}]^2\sqrt[4]{4.4\times12} = 1.79 \text{ cub. ft. per sec.}$$ The tube must be at least 3 times as long as wide, to be filled. 509. Conical Diverging, and Converging, Short Tubes.--With conical convergent tubes, as at A, Fig. 565, with inner edges not rounded, D'Aubuisson and Castel found by experiment values of the coefficient of velocity, φ, and of that of efflux μ [from which the coefficient of contraction, C = μ + φ, may be computed.] for tubes 1.55 centimeters wide at the narrow end, and 4.00 centimeters long, under a head of h = 3 metres, and with different angles of convergence. By angle of convergence was meant the angle between the sides CE and DE, Fig. 565. In the following table will be found some values of x and φ founded on these experiments, for use in the formula $$v_m = \phi \sqrt{2gk} \quad \text{and} \quad Q = \mu F \sqrt{2gk};$$ in which $F$ denotes the area of the outlet orifice $EB$. A diagram showing a conical diverging tube with a converging section. Fig. 565. CONICAL SHORT PIPES. 693
TABLE G (CONICAL CONVERGING TUBE).
Angle of convergence 8° 10' 8° 10' 90' 18° 90' 18° 90' 80° 45°
$\mu$ = .965 .965 .965 .965 .965 .965 .965
$\phi$ = .964 .964 .964 .964 .964 .964 .964
Evidently $\mu$ is a maximum for $18^{\circ}$. With a conically divergent tube at $MN$, having the internal diameter $MO = .025$ metre, the internal diam. $NP = .032$ metre, and the angle between $MN$ and $PN$ $= 18^{\circ}$, Weibull found that with $F = .027$ square metres, $\frac{V_2}{V_1} = \sqrt{\frac{F}{\pi d}}$ (where $F =$ area of outlet section $NP$) $\mu$ should be $=.553$; the great loss of velocity as compared with $\sqrt{\frac{d}{\pi}}$ being due to the eddying in the re-expansion from the contracted section at $M$ (corners not rounded), as occurs also in Fig. 549. The jet was much troubled and pulsed violently. When the angle of divergence is too great, or the head $h$ too large, or if the tube is not wet by the water, efflux with the tube filled cannot be maintained, the flow then taking place as in Fig. 550. Venturi and Eytelwein experimented with a conically divergent tube (called now "Venturi's tube"), with rounded entrance to conform to the shape of the contracted vein, as in Fig. 566, having a diameter of one inch at $m$ (narrowest part), where the sectional area is $A = .0783$. At $m$, where $V_2 = 1.80$ inches at $m$ (outlet), where area $= F'$; the length being 8 ins., and the angle of convergence $5^{\circ} 9'$. With $Q = \mu F' V_2 \frac{d}{\pi h}$ they found $\mu = .483$. Hence 24 times as much water was discharged as would have flowed out under the same head through an orifice in thin plate with area $= F' =$ the smallest section of the divergent tube, and 1.9 times as much as through a short pipe of section $= F'$. A similar calculation shows that the velocity at $m'$ when $\mu = .483$, and $\phi = .964$, showed that the pressure at $m'$ was much less than one atmosphere. Fig. 550. 694 MECHANICS OF ENGINEERING. Mr. J. B. Francis also experimented with Venturi's tube (see "Lowell Hydraulic Experiments"). See also p. 389 of vol. 6 of the Journal of Engineering Societies, for experiments with diverging short tubes discharging under water. The highest coefficient ($\mu$) obtained by Mr. Francis was 0.782. 509a. New Forms of the Venturi Tube.—The statement made in § 507, in connection with Fig. 562, was based on purely theoretic grounds, but has recently (Dec. 1888) been completely verified by experiments conducted in the hydraulic laboratory of the Massachusetts Institute of Technology, at Cambridge, Mass., under the direction of Professor W. H. Emmons, of Harvard University. Three short tubes of circular section, each 3 in. in length and 1 in. in internal diameter at both ends, were experimented with, under heads of 2 ft. and 4 ft. Call them A, B and C. A was an ordinary straight tube as in Fig. 561; the longitudinal section of B was like that in Fig. 562, the narrowest diameter being 0.80 in. (see § 495; $0.8^{\prime} = 0.64$); while C was somewhat like that in Fig. 566, being formed like B up to the narrowest part (diameter 0.80 in.), and then made gradually divergent to the discharging end. The results of the experiments are given in the following table:
Name of Tube Head. Number of trials. Range of Values of $\mu$. Average value of $\mu$.
A A = 2 ft. 4 From 0.90 to 0.98 0.914
A A = 4 ft. 4 From 0.818 to 0.98 0.931
B A = 2 ft. 5 * * * 0.975 to 0.995 0.983
B A = 4 ft. 4 * * * 0.981 to 0.998 0.992
C A = 2 ft. 5 * * * 0.990 to 0.919 0.901
C A = 4 ft. 4 * * * 0.908 to 0.983 0.914
The fact that B discharges more than A is very noticeable, while the superiority of C to B, though evident, is not nearly so great as that of B to A, showing that in order to increase the discharge of an (originally) straight tube (by encroaching on the passage-way) it is of more importance to fill up with solid substance the space around the contracted vein than to make the transition from the narrow section to the discharging end very gradual. * See Journal of the Franklin Inst., for April, 1889. Description: A diagram showing three different forms of Venturi tubes used for hydraulic experiments. "FLUID FRICTION." 695 510. "Fluid Friction."—By experimenting with the flow of water in glass pipes inserted in the side of a tank, Prof. Reynolds of England has found that the flow goes on in parallel filaments for only a few feet from the entrance of the tube, and that then the liquid particles begin to intermingle and cross each other's paths in the most intricate manner. To render this more vividly, I visited a laboratory where colored liquid at the inlet of the pipe and observed its further motion, and found that the greater the velocity the nearer to the inlet was the point where the breaking up of the parallelism of flow began. The hypothesis of laminated flow is, nevertheless, the simplest theoretical basis for establishing practical formulae, and the resistance offered by pipes to the flow of liquids in them will therefore be attributed to the friction of the edges of the lamina against the inner surface of the pipe. The amount of this resistance (often called skin-friction) for a given extent of rubbing surface is by experiment found— 1. To be independent of the pressure between the liquid and the solid; 2. To vary nearly with the square of the relative velocity; 3. To vary directly with the amount of rubbing surface; 4. To vary directly with the heaviness ($y$, § 408) of the liquid. Hence for a given velocity $v$, a given rubbing surface of area $S$, and a liquid of heaviness $y$, we may write $$\text{Amount of friction (force)} = fSy \frac{v^2}{2g}, \quad \ldots \ldots \ldots (1)$$ in which $f$ is an abstract number called the coefficient of fluid friction, to be determined by experiment. For a given liquid, given character (roughness) of surface, and small range of velocities it is approximately constant. The object of introducing $\frac{v^2}{2g}$ is not only because $\frac{v^2}{2g}$ is a familiar and useful function of $v$, but that $v^2 - 2g$ is a height, or distance, and therefore the product of $S$ (an area) by $v^2 - 2g$ is a volume, and this volume multiplied by $y$ gives the weight of an ideal prism of A diagram showing a cross-section of a pipe with water flowing through it. 096 MECHANICS OF ENGINEERING. the liquid; hence $S \frac{v}{2g} r$ is a force and $f$ must be an abstract number and therefore the same in all systems of units, in any given case or experiment. In his experiments at Torquay, England, the late Mr. Froude found the following values for $f$, the liquid being salt water, while the rigid surface was on two sides of a thin straight wooden bar $A$, of an inch thick and 19 inches long, polished or prepared in various ways, and drawn edgewise through the water at a constant velocity, the total resistance being measured by a dynamometer. 511. Mr. Froude's Results.—(Condensed.) [The velocity was the same = 10 ft. per sec. in each of the following cases. For other velocities the resistance was found to vary nearly as the square of the velocity, the index of the power varying from 1.8 to 2.16.]
Character of Surface. Value of $f$ (from eq. (3), § 18).
When the board was
2 ft. long 8 ft. long 16 ft. long
Varnish $J =$ 0.064 0.0028 0.0098 0.0055
Pearlinae 0.008 0.001 0.007
Tufalite 0.0087 0.003 0.015 0.025
Calcio. 0.007 0.063 0.033 0.067
Fine Sand. 0.069 0.254 0.169 0.499
Medium Sand. 0.069 0.262 0.168 0.499
Coarse Sand. 0.119 0.771 0.569
N.B. These numbers multiplied by 10 also give the same fractional resistance in the case of a pipe of 1 foot diameter, but with a cross-section equal to that of a square of area 64 sq. ft., per cubic foot, to cancel the $J$ = 64 sq ft per sec. of eq. (3) of this section.
For use in formula bearing on flow in pipes, $f$ is best determined directly by experiments of that very nature, the results of which will be given as soon as the proper formulae have been established. 512. Bernoulli's Theorem for Steady Flow, with Friction.—[The student will now re-read the first part of § 492, as far as eq. (1).] Considering free any laminar of fluid, Fig. 567, (according to the subdivision of the stream agreed upon in § 492 referred BERNOULLI'S THEOREM WITH FRICTION. 697 to,) the frictions on the edges are the only additional forces as compared with the system in Fig. 534. Let $w$ denote the length of the *rected perimeter* of the base of this lamina (in case of a pipe running full, as we here postulate), then $w$ is of course the *whole perimeter*, but in the case of an open channel or canal, $w$ is only a portion of the whole perimeter of the cross-section). Then, since the area of rubbing surface at the edge is $S = w d e$, the total friction for the lamina is [by eq. (1), § 510] $= f w y (e^2 + 2 g) d e$. Hence from $w e = (\tan . \sec .) \times d e$, and from $(\tan . \sec .) = [\frac{2}{g}(\tan . \cos . \text{comp. of acting forces})]$ $\times$ (mass of lamina), we have $$Fp - F(p + dp) + Fyde \cos \phi - f w y \frac{v^2}{g} de = Fyde + g \cdot de... (a)$$ As in § 492, so here, considering the simultaneous advance of all the lamina lying between any two sections $m$ and $n$ during the small time $dt$, putting $de = ds$, and $de \cos \phi = -ds$ (see Fig. 568), we have, for any one lamina, $$\frac{1}{g} v dv + \frac{1}{g} dp + ds = -f w \frac{v^2}{g} dy... (1)$$ Now conceive an infinite number of equations to be formed like eq. (1), one for each lamina between $m$ and $m$, for the same $dt$, viz., a $de$ of such length that each lamina at the end of $dt$ will occupy the same position, and acquire the same value of $v$, and so that the lamina had in its front had at the beginning of the $dt$ (this is the characteristic of a steady flow). Adding up A diagram showing a pipe with water flowing through it, illustrating Bernoulli's theorem with friction. A diagram showing a pipe with water flowing through it, illustrating Bernoulli's theorem with friction. A diagram showing a pipe with water flowing through it, illustrating Bernoulli's theorem with friction. A diagram showing a pipe with water flowing through it, illustrating Bernoulli's theorem with friction. 698 MECHANICS OF ENGINEERING. the corresponding terms of all these equations, we have (re- membering that for a liquid $y$ is the same in all laminae), $$\frac{1}{g}\int_{a}^{b}v\,ds+\frac{1}{y^2}\int_{a}^{b}dp+\int_{a}^{b}zds=-\frac{f}{2g}\cdot\int_{a}^{b}\frac{w}{F}v^2ds;\tag{9}$$ i.e., after transposition and writing $R$ for $F+v$, for brevity, $$\frac{u^2}{2g}+\frac{P_s}{y}+s_0=\frac{u^2}{2g}+\frac{P_s}{y}+s_0-\frac{f}{2g}\int_{a}^{b}\frac{v^2}{R}ds.\tag{9'}$$ This is Bernoulli's Theorem for steady flow of a liquid in a pipe of slightly varying sectional area $F$, and internal perim- eter $w$, taking into account no resistances or friction, except the "skin-friction," or "fluid-friction" of the liquid and sides of the pipe. Resistances due to the internal friction of eddying occasioned by sudden enlargements of the cross-section of the pipe, by elbows, sharp curves, valve-gates, etc., will be mentioned later. The negative term on the right in (3) is of course a height or head (one dimension of length), as all the other terms are such, and since it is the amount by which the sum of the three heads (viz., velocity-head, pressure-head, and potential head) at $m$, the down-stream position, lacks of being equal to the sum of the corresponding heads at $n$, the up-stream position or section, we may call it the "Loss of Head" due to skin-friction between $a$ and $m$; also called *friction-head*, or resistance-head, or height of loss. The quantity $R=F+v$ is sectional-area + wetted-perimeter, is an imaginary distance or length called the Hydraulic Mean Radius, or Hydraulic Mean Depth, or simply hydraulic radius of the section. For a circular pipe of diameter $d$, $$R=\frac{\pi d^2}{4\pi d^2}=d;\tag{10'}$$ while for a pipe of rectangular section, $$R=\frac{ab}{2(a+b)}.\tag{10'''}$$ **FRICITION IN PIPES.** 513. Problems involving Friction-heads; and Examples of Bernoulli's Theorem with Friction. **PROBLEM I.**—Let the portion of pipe between $n$ and $m$ be level, and of uniform circular section and diameter = $d$. The jet at $m$ discharges into the air, and has the same sectional area, $F = \pi d^2$, as the pipe from which the pressure-head at $m$ is $$P_m = b = 34 \text{ feet (for water), and the velocity head at } m \text{ is that at } n, \text{ since } v_n = v_m.$$ The height of the water column in the open piezometer at $n$ is noted, and is $y_n$ (so that the pressure-head at $n$ is $\frac{y_n}{g} = y_n + h$); while the length of pipe from $n$ to $m$ is $l$. Knowing $l$, $d$, $y_n$, and having measured the volume $Q$ of flow per unit of time, it is required to find the form of the friction-head and the value of $f$. From $$F_n v_n = Q_1 \text{ or } \frac{\pi d^2 v_n}{g} = Q_1 \dots \dots (1)$$ $v_n$ becomes known. Also, $v_n$ is known to be $= v_m$, and the velocity at each $ds$ is $v = v_m$, since $F(sections\ area)$ is constant along the pipe, and $F_0 = Q$. The hydraulic radius is $$R = \frac{l}{d}; \dots \dots (2)$$ the same for all the $ds$'s between $n$ and $m$. Substituting in eq. (3) of § 312, with the horizontal axis of the pipe as a datum for potential heads, we have $$\frac{v_m^2}{2g} + b + 0 = \frac{v_n^2}{2g} + y_n + b + 0 - \frac{l}{d^2} \cdot \frac{v_m^2}{2g} ds; \dots (3)$$ i.e., since $\int ds = l =$ length of pipe from $n$ to $m$, the friction-head for a pipe of length $l$, and uniform circular section of diameter = $d$, reduces to the form 700 MECHANICS OF ENGINEERING. Friction-head = 4f $\frac{l}{d}$ $\frac{v^2}{2g}$ ; . . . (6) where $v$ = velocity of water in the pipe, being in this case also $= v_{w}$ and $= v_{s}$. Hence this friction-head varies directly as the length and as the square of the velocity, and inversely as the diameter; also directly as the coefficient $f$. From (3), then, we derive (for this particular problem) Piezometer-height at $n = y_n = 4f \frac{l}{d} \frac{v^2}{2g}$ ; . . . (5) i.e., the open piezometer-height at $n$ is equal to the loss of head (all of which is friction-head here) sustained between $n$ and the mouth of the pipe. (Pipe horizontal). EXAMPLE. Required the value of $f$, knowing that $d = 3$ in., $y_1$ (by observation) = 10.4 ft., and $Q = 0.1960$ cub. ft. per sec., while $l = 400$ ft. ($n$ to $m$). From eq. (1) we find, in ft.-lb.-sec. system, the velocity in the pipe to be $v = \frac{4Q}{\pi d^2} = \frac{4 \times 0.1960}{\pi \times 3^2} = 4.0$ ft. per sec; then, using eq. (5), we determine $f$ to be $f = \frac{9py_1d}{4Ql} = \frac{9 \times 82.9 \times 1 \times 10.4}{4 \times 400 \times 3^2} = 0.0065.$ PROBLEM II. Hydraulic Accumulator—Fig. 570. Let the area $F_2$ of the piston on the left be quite large compared with that of the pipes and nozzle. The cylinder contains a fricton- **FRICITION-HEAD IN PIPES.** 701 less weighted piston, producing (so long as its downward slow motion is uniform) a fluid pressure on its lower face of an intensity $p_n = [G + F_n p_a] - F_n$ per unit area ($p_n = \text{one stmos}$). Hence the pressure-head at $n$ is $$\frac{p_n}{y} = \frac{G}{F_n} + b_n \quad \ldots \ldots \quad (6)$$ where $G = \text{load on piston}$. The jet has a section $a = F_a$, = that of the small straight nozzle (no contraction). The junctions of the pipes with each other, and with the cylinder and nozzle, are all smoothly rounded; hence the only losses of head in steady flow between $a$ and $m$ are the friction-heads in the two long pipes, neglecting that in the short nozzle. These friction-heads may be of the form in eq. (4), and will involve the velocities $v_a$ and $v_m$, respectively in these pipes (supposed running full). $v_a$ and $v_m$ may be unknown at the outset, as here. Known to us are the lengths and diameters of the pipes, we are required to find the velocity $v_a$ of the jet, flowing into the air, and the volume of flow, $Q$, per unit of time, assuming $f$ to be known and to be the same in both pipes (not strictly true). Let the lengths and diameters be denoted as in Fig. 570, their sectional areas $F_a$ and $F_m$, the unknown velocities in them $v_a$ and $v_m$. From the equation of continuity [eq. (3), § 490], we have $$v_a = \frac{F_an}{F_a} \quad \text{and} \quad v_m = \frac{F_mn}{F_m} \quad \ldots \ldots \quad (7)$$ To find $v_m$, we apply Bernoulli's Theorem (with friction), eq. (8), § 512, taking the down-stream position $m$ in the jet close to the nozzle, and the up-stream position $n$ just under the piston in the cylinder where the velocity $v_a$ is practically nothing. Then with $m$ as datum plane we have $$\frac{u^2_m}{2g} + b + 0 + G + F_m n - 4f \frac{l}{d_m} u^2_m = 4f \frac{l}{d_a} u^2_a \quad \ldots \ldots \quad (8)$$ 702 MECHANICS OF ENGINEERING. Apparently (8) contains three unknown quantities, $v$, $\tau$, and $v_1$; but from eqs. (7) $v$ and $v_1$ can be expressed in terms of $w$, whence [see also eq. (6)]: $$\frac{w}{2} \left[ 1 + 4f \frac{F_2'}{R} + 4f' \frac{F_2''}{R} \right] = A + \frac{G}{F_2'}; \quad (9)$$ or, finally, $$w = \sqrt{\frac{3g(h - G)}{1 + 4f \frac{F_2'}{R} + 4f' \frac{F_2''}{R}}} \cdot \quad (10)$$ and hence we have also $$Q = F_2 w \cdot \quad \quad (11)$$ **EXAMPLE.—If we replace the force $G$ of this problem by the thrust $P$ exerted along the pump-platton of a steam fire-engine, we may treat the foregoing as a close approximation to the practical problem of such an apparatus, the pipes being connective straight lengths of hose. In which (for the probable values of $h$, $G$, and $P$ given in § 507), we have [see, e.g., "Steam Engineering," by Geo. Ellis, Springfield, Mass.),* (Strictly, $f$ varies somewhat with the velocity; see § 517.) Let $P = 12000$ lbs., and the piston-area at $n = F_2 = 78$ sq. in. = $a$ sq. ft. Also, let $h = 20$ ft., and the dimensions of the hose be as follows: $$d_1 = 8 \text{ in.}, \quad d_2 = 2 \text{ in.}, \quad d_a (\text{of nozzle}) = 1 \text{ in.};$$ $$l_1 = 400 \text{ ft.}, \quad l_2 = 500 \text{ ft.}$$ With the foot-pound-second system of units, we now have [eq. (10)] $$w = \sqrt{\frac{2 \times 32.2}{1 + 4f \times .0075}} = \sqrt{\frac{64.4}{1 + .03}} = \sqrt{\frac{64.4}{1.03}}.$$* * See § 528 **FRICTION-HEAD IN PIPES.** 708 $$= \sqrt{\frac{3 \times 39.2 \times 404}{1 + 0.39 + 5.68}}$$ I.e., $v_m = 60.0$ ft. per sec. If this jet were directed vertically upward it should theoretically attain a height of $\frac{v_m^2}{2g} = nearly$ 56 feet, but the resistance of the air would reduce this to about 40 or 45 ft. We have further, from eq. (1), $$Q = F_{av} v_m = \frac{\pi(1)^2}{4(12)} \times 60.0 = 3.37 \text{ cub. ft. per sec.}$$ If there were no resistance in the hose we should have, from § 497a, $$v_m = \sqrt{2g\left(\frac{P}{F_{av}} + h\right)} = \sqrt{2 \times 39.2 \times 404} = 161.3 \text{ ft. per sec.}$$ §13a. Influence of Changes of Temperature.—Although Poisson and Hagen found that with glass tubes of very small diameter the flow of water was increased threefold by a rise of temperature of one degree, yet it is understood that with common pipes the rate of flow is appreciably affected by ordinary fluctuations of temperature; at any rate, experiments of sufficient precision are wanting, as regards such an influence. See Mr. Hamilton Smith's "Hydraulics," p. 16, where he says: "Changes by variation in $T$ (temperature) will probably only be appreciable with small orifices, or with very low heads for orifices or weirs." §14. Loss of Head in Orifices and Short Pipes.—So long as the steady flow between two localities $a$ and $b$ takes place in a pipe having no abrupt enlargement or diminution of section, nor sharp curves, bends, or elbows, the loss of head may be ascribed solely to the surface action (or "skin-friction") between water and pipe; but the introduction of any of the above-mentioned features occasions eddying and internal disturbance, and friction (and consequent heat) thereby causing further deviations 704 MECHANICS OF ENGINEERING. from Bernoulli's Theorem; i.e., additional losses of head, or heights of resistance. From the analogy of the form of a friction-head in a long pipe [eq. (4), § 513], we may assume that any of the above heights of resistance is proportional to the square of the velocity, and may therefore always be written in the form $$\left\{\begin{array}{l}{\text{Loss of Head due to any}}\\ {\text{cause except skin-friction}}\end{array}\right\} = \zeta \frac{v^2}{gq}, \quad \ldots \ldots \ldots (1)$$ in which $v$ is the velocity of the water in the pipe at the section where the resistance occurs; or, if, on account of a abrupt enlargement of the stream-section, there is a corresponding diminution of velocity, then $v$ is always to denote this diminished velocity (i.e., in the down-stream section). This velocity $v$ is often an unknown at the outset. $\zeta$, corresponding to the abstract factor $\frac{4}{D}$ in the height of resistance due to skin-friction [eq. (4), § 513], is an abstract number called the Coefficient of Resistance, to be determined experimentally; or computed theoretically, where possible. Roughly speaking, it is independent of the velocity, for a given fitting, casing, pipe-joint, elbow, bend, valve-gate at a definite opening, etc., etc. 515. Heights of Resistance (or Losses of Head) Occasioned by Short Cylindrical Tubes.—When dealing with short tubes discharging into the air, in § 507, deviations from Bernoulli's Theorem were made good by using a coefficient of velocity $\phi$, dependent on experiment. This device answered every purpose for the simple circumstances of the case, as well as for simply orifices. But the great variety of possible designs of a component pipe (valve, elbow, fitting, etc.) and the various shapes of cross-sections (etc.) renders it almost impossible, in such a pipe, to provide for deviations from Bernoulli's Theorem by a single coefficient of velocity (velocity of jet, that is) for the pipe as a whole, since new experiments would be needed for each new design of pipe. Hence the great utility of the conception of "loss of head," one for each source of resistance. LOSSES OF HEAD IN A SHORT PIPE. 705 If a long pipe issues from the plane side of a reservoir and the corners of the junction are not rounded [see Fig. 571], we shall need an expression for the loss of head at the entrance, $E$, as well as that $$\left[ = 4 \frac{f}{d} \frac{l}{2g} \right]$$ due to the skin-friction in the pipe. But, whatever the velocity, $v$, in the pipe proves to be, influenced as it is both by the entrance loss of head and the skin-friction head (in applying Bernoulli's Theorem), the loss of head at $E$, viz., $\zeta_{E} \frac{v^2}{2g}$ will be just the same as if efflux took place through enough of the pipe at $E$ to constitute a "short pipe," discharging into the air, under some head $h$ (different from $h'$ of Fig. 571) sufficient to produce the same velocity $v$. But in that case we should have $$v = \phi \sqrt{2gh}, \quad \text{or} \quad \frac{v^2}{2g} = \phi^2 h. \quad . . . . . . . . (1)$$ (See §§ 507 and 508, $\phi$ being the coefficient of velocity, and $h$ the head, in the cases mentioned in those articles.) We, therefore, apply Bernoulli's Theorem to the cases of those articles (e.g. Figs. 540 and 564), in order to determine the loss of head due to the short pipe, and obtain (with $m$ as datum level for potential heads) $$\frac{vm}{2g} + b + 0 = 0 + b + h - \zeta_{E} \frac{v^2}{2g}. \quad . . . . . . . . (2)$$ Now the $v$ of eq. (2) is equal to the $v_m$ of the figures referred to, and $\zeta_E$ is a coefficient of resistance for the short pipe, and we now desire its value. Substituting for $$\frac{v^2}{2g} = \frac{vm}{2g}$$ we get $$b + h - \zeta_E \frac{vm}{2g} = 0.$$ 706 MECHANICS OF ENGINEERING. its value $\varphi^{\prime}h$ from eq. (1), we have $$\zeta_{e} = \frac{1}{\varphi^{\prime}} - 1 \quad \ldots \quad \ldots \quad \ldots \quad (5)$$ Hence when $\alpha = 90^\circ$ (i.e., the pipe is $\gamma$ to the inner reservoir surface), we derive $$\zeta_{e} = \frac{1}{\varphi^{\prime}} - 1 = \frac{1}{(0.816)^{\alpha}} - 1 = 0.505; \quad . . . \quad [\alpha = 90^\circ] \quad . . .$$ and similarly, for other values of $\alpha$ (taking $\varphi$ from the table, § 508), we compute the following values of $\zeta_{e}$ (corners not rounded) for use in the expression for “loss of head,” $c_{e}^{2} \cdot \frac{v}{2g}$:
For $\alpha = 90^\circ$ $\zeta_{e} = 0.505$ $\zeta_{e} = 0.665$ $\zeta_{e} = 0.635$ $\zeta_{e} = 0.718$ $\zeta_{e} = 0.784$ $\zeta_{e} = 0.870$ $\zeta_{e} = 0.965$
For $\alpha = 70^\circ$ $\zeta_{e} = 0.707$ $\zeta_{e} = 0.787$ $\zeta_{e} = 0.833$ $\zeta_{e} = 0.871$ $\zeta_{e} = 0.914$ $\zeta_{e} = 0.965$ $\zeta_{e} = 1.025$
For $\alpha = 60^\circ$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$ $\zeta_{e} = 0.787$ $\zeta_{e} = 0.833$ $\zeta_{e} = 0.871$ $\zeta_{e} = 0.914$ $\zeta_{e} = 0.965$
For $\alpha = 50^\circ$ $\zeta_{e} = 0.595$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$ $\zeta_{e} = 0.787$ $\zeta_{e} = 0.833$ $\zeta_{e} = 0.871$ $\zeta_{e} = 0.914$
For $\alpha = 40^\circ$ $\zeta_{e} = 0.525$ $\zeta_{e} = 0.595$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$ $\zeta_{e} = 0.787$ $\zeta_{e} = 0.833$ $\zeta_{e} = 0.871$
For $\alpha = 30^\circ$ $\zeta_{e} = 0.444$ $\zeta_{e} = 0.525$ $\zeta_{e} = 0.595$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$ $\zeta_{e} = 0.787$ $\zeta_{e} = 0.833$
For $\alpha = 20^\circ$ $\zeta_{e} = 0.344$ $\zeta_{e} = 0.444$ $\zeta_{e} = 0.525$ $\zeta_{e} = 0.595$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$ $\zeta_{e} = 0.787$
For $\alpha = 10^\circ$ $\zeta_{e} = 0.222$ $\zeta_{e} = 0.344$ $\zeta_{e} = 0.444$ $\zeta_{e} = 0.525$ $\zeta_{e} = 0.595$ $\zeta_{e} = 0.666$ $\zeta_{e} = 0.733$
For $\alpha < 10^\circ$, etc.
From eq. (4) we see that the loss of head at the entrance of the pipe, corners not rounded, with $\alpha=90^\circ$, is about one half (.5/2) of the height due to the velocity $v_1^n$, in that part of the pipe ($v_1^n\) being the same all along the pipe if cylindrical). The value of $v_1^n$, itself, Fig.\,571, depends on all the features of the design from reservoir to nozzle.\ See §~518. If the corners at $E'$ are properly rounded, the entrance loss of head may practically be done away with; still, if $v_1^n$is quite small (as it may frequently be, from large losses of head farther down stream), the saving thus secured, while helping to increase $v_1^n$slightly (and thus the saving itself), is insignifi- cant. §~518 General Form of Bernoulli's Theorem, considering all Losses of Head. In view of preceding explanations and assumptions, we may write a general and final form Bernoulli's Theorem for steady flow from an up-stream position $n$a down-stream position $m$, as follows: $$c_n^2 + \frac{P_n}{y'} + z_n + c_m^2 + z_m - \left(\text{all losses of head occurring between } n \text{ and } m \right) . (B_2)$$ **LOSSES OF HEAD IN GENERAL.** 707 Each loss of head (or height of resistance) will be of the form $$\zeta = \frac{v^2}{2g}$$ (except skin-friction head in long pipes, viz., $4f \frac{v^2}{d}$), the $v$ in each case being the velocity, known or unknown, in that part of the pipe where the resistance occurs (and hence is not necessarily equal to $u_a$ or $u_b$). 517. The Coefficient, $f$, for Friction of Water in Pipes—See eq. (1), § 510. Experiments have been made by Weibach, Eitelwein, Darby, Bosanet, Prony, Dubuat, Fanning, and others, to determine $f$ in cylindrical pipes of various materials (tin, glass, zinc, lead, brass, cast and wrought iron) of diameters from $\frac{3}{8}$ inch up to 36 inches. In general, the following deductions may be made from these experiments: 1st. $f$ decreases when the velocity increases; e.g., in one case with the same pipe $f$ was = .0070 for $v = 2'$ per sec., while $f$ was = .0036 for $v = 20'$ per sec. 2ndly. $f$ decreases slightly as the diameter increases (other things being equal); e.g., in one experiment $f$ was = .0069 in a 3-in. pipe, while for the same velocity $f$ was = .0084 in a 6-in. pipe. 3rdly. The condition of the interior surface of the pipe affects the value of $f$, which is larger with increased roughness of pipe. Thua, Darby found, with a foul iron pipe with $d = 10$ in. and velo. = 3.67 ft. per sec., the value 0113 for $f$; whereas Fanning (see p. 338 of his "Water-supply Engineering"), with a cement-lined pipe and velocity of 3.74 ft. per sec. and $d = 20$ inches, obtained $f = .0052$. Weibach, finding the first relation very prominent, proposed the formula $$f = 0.00359 + \frac{.00499}{V_0 (\text{in ft. per sec.)}}$$ when the velocities are great, while Darby, taking into account both the 1st and 3rd relations above, writes (see p. 585, Rankine's Applied Mechanics) A small image of a water flow diagram. 708 MECHANICS OF ENGINEERING. $$f = \frac{0.045}{\left(1 + \frac{1}{9} \times \text{diam. in ft}\right)} + \frac{0.001}{0 \text{ ft per sec.}} \left(1 + \frac{1}{18} \times \text{diam. in ft}\right)$$ For practical purposes, Mr. J. T. Fanning has recommended, and arranged in an extensive table (pp. 349-346 of his book just mentioned), values of $f$ for clean iron pipe, of diameters from $\frac{3}{4}$ inch to 96 inches, and for velocities of 0.1 ft. to 30 ft. per second, for the pipe opposite to that shown, inserted with Mr. Fanning's permission, for use in solving numerical problems. In obtaining $f$ for slightly tuberculated and for foul pipes, the recommendations of Mr. Fanning seem to justify the following rules: For slightly tuberculated pipes of diam. = $\frac{3}{4}$ ft., 1 ft., 9 ft., 4 ft. we should add 23% 34% 10% 13% and for foul pipes of same size............ 72% 60% 38% 25% of the $f$ for clean pipes, to itself. For example, if $f = .007$ for a certain $\frac{3}{4}$ ft. pipe when clean, with velocity = 0.04 ft. per sec., we have $f' = .007 \times 1.72 = .01204$ when it is foul. For first approximations a mean value of $f = .006$ may be employed, since in some problems sufficient data may not be known in advance to enable us to take $f'$ from the table. **EXAMPLE—Fig. 572.** In the steady pumping of crude petroleum weighing $y = .55$ lbs. per gallon through a six-inch pipe 30 miles long, to a station 700 ft. higher than the pump, it is found that the pressure in the pump cylinder at $n$, necessary to keep up a velocity of 4.4 ft. per sec., is equal to 1,000 lbs. per sq. inch. Required the coefficient $f$ of the pipe. As all losses except the friction-head in the pipe are insignificant, the latter only will be considered. The velocity-head at $n$ may A diagram showing a pipeline with a pump and a gauge. Fig. 572 lbs. per sq. inch. Required the coefficient $f$ of the pipe. As all losses except the friction-head in the pipe are insignificant, the latter only will be considered. The velocity-head at $n$ may
TABLE OF VALUES OF f. 709
P (mm)                                                 f (mm)                                                                                                       & f (mm) &
1.0 0.0085 0.0086
1.5 0.0086 0.0087
2.0 0.0087 0.0088
2.5 0.0088 0.0089
3.0 0.0089 0.0091
3.5 0.0091 0.0092
4.0 0.0092 0.0094
4.5 0.0094 0.0096
5.0 0.0096 0.0112
5.5 0.0112 0.0114
6.0 0.0114 0.0116
6.5 0.0116 0.0118
7.0 0.0118 0.0122
7.5 0.0122 0.0124
8.0 0.0124 0.0126
8.5 0.0126 0.0128
9.0 0.0128
710 MECHANICS OF ENGINEERING. be put $=0$; the jet at $m$ being of the same size as the pipe, the velocity in the pipe is $=v_m$, and therefore $v_n = 4.4$ ft. per sec. Notice that $m$, the down-stream section, is at a higher level than $n$. From Bernoulli's Theorem, § 516, we have, with $s$ as a datum level, $$\frac{v_m^2}{2g} + h + b = \frac{F_p}{y} + 0 - \frac{4fL}{d}\frac{v^2}{2g}, \quad (1)$$ Using the ft., lb., and sec., we have $$h = 700 \text{ ft.}, \quad v_m^2 + 2g = 0.30 \text{ ft.},$$ while $$b = \frac{14.7 \times 144}{55} = 88.47 \text{ ft.}, \quad \frac{F_p}{y} = \frac{1000 \times 144}{55} = 2618 \text{ ft.}$$ Hence, in eq. (1), $$0.30 + 88.4 + 700 = 2618 - \frac{4fL}{d} \times \frac{30 \times 5280}{64.4}.$$ Solving for $f$, we have $f = .00485$ (whereas for water, with $v = 4.4$ ft. per sec. and $d = \frac{1}{2}$ ft., the table, p. 709, gives $f = .00601$). If the $y$ of the petrolum had been 50 lbs. per cubic foot, instead of 55, we would have obtained $\frac{F_p}{y} = 2880$ feet and $f = .0058$. 518. Flow through a Long Straight Cylindrical Pipe, including both friction-head and entrance loss of head (corners not rounded); reservoir large. Fig. 573. The jet issues directly from the end of the pipe, in parallel filaments, into the air and therefore has no resistance to flow; hence, also, $v_n$ of the jet $= v$ in the pipe (which is assumed to be running full), and is A diagram showing a long straight cylindrical pipe with a jet issuing from its end into the air.Fig. 573 COEFFICIENT OF LIQUID FRICTION. 711 therefore the velocity to be used in the loss of head $c_{w} \frac{v^2}{2g}$ at the entrance $E$ (§ 815). Taking $m$ and $n$ as in figure and applying Bernoulli's Theorem (§ 476), with $m$ as datum level for the potential heads $\sigma_m$ and $\sigma_n$, we have $$\frac{\sigma_n + b + 0 = 0 + b + h - c_w \frac{v^2}{2g}}{2g} = 4f \frac{l}{d} \frac{v^2}{2g} \cdot \quad (1)$$ Three different problems may now be solved: First, required the head $h$ to keep up a flow of given volume $Q$ per unit of time in a pipe of given length $l$ and diameter $d$. From the equation of continuity we have $$Q = F_a v_m = \frac{4Q}{\pi d^2} v_m;$$ i.e., veloc. of jet, which = veloc. in pipe, $= v_m = \frac{4Q}{\pi d^2} \cdot \quad (2)$$ Having found $v_m = v$, from (2), we obtain from (1) the required $h$, thus: $$h = \frac{v^2}{2g} \left[1 + c_w + 4f \frac{l}{d}\right] \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)$$ Now $c_w = 0.505$ if $\alpha = 90^\circ$ (see § 815), while $f$ may be taken from the table, § 817, for the given diameter and computed velocity ($v_m = v$, found in (2)), if the pipe is clean; if not clean, see end of § 817, for slightly tuberculated and for four pipes. Secondly. Given the head $h$, and the length $l$ and diameter $d$ of pipe, required the velocity in the pipe, viz., $v = v_m$, that of jet; also the volume delivered per unit of time, $Q$. Solving eq. (1) for $v_m$, we have $$v_m = \sqrt{\frac{1}{1 + c_w + 4f \frac{l}{d}}} \cdot \quad (4)$$ 712 MECHANICS OF ENGINEERING. whence $Q$ becomes known, since $$Q = \frac{4}{\pi} d v_{w} \quad \quad \quad (5)$$ [Note.—The first radical in (4) might for brevity be called a coefficient of velocity, $\phi$, for this case. Since the jet has the same diameter as the pipe, this radical may also be called a coefficient of efflux.] Since in (4) $f$ depends on the unknown $v$ as well as on the known $d$, we must first put $f = .006$ for a first approximation for $v$, and find a corresponding value for $f$ and substitute again; and so on. Thirdly, knowing the length of pipe and the head $h$, we wish to find the proper diameter $d$ for the pipe to deliver a given volume $Q$ of water per unit of time. Now $$v_1 = v_{w} = \frac{Q}{4 \sqrt{\pi h}} \quad \quad \quad (6)$$ which substituted in (1) gives $$2gh = \left(\frac{4Q}{\pi}\right)^{\frac{1}{2}}\left[1 + c_2 + 4f^{\frac{1}{2}}\right] = \left(\frac{4Q}{\pi}\right)\left[\frac{1+c_2}{d} + 4f^{\frac{1}{2}}\right];$$ that is, $$2ghd^2 = \left(\frac{4Q}{\pi}\right)\left[(1+c_2)d + 4f^{\frac{1}{2}}\right];$$ $$\therefore d = \sqrt{\frac{(1+c_2)d + 4f^{\frac{1}{2}}}{2gh}\left(\frac{4Q}{\pi}\right)} \quad \quad \quad (7)$$ As the radical contains $d$, we first assume a value for $d$, with $f = .006$, and substitute in (7). With the approximate value of $d$ thus obtained, we substitute again with a new value for $f$ based on an approximate $v$ from eq. (6) (with $d$ its first approximation), and so on until the value for $d$ is derived; and so on. (Treutwine's Pocket-book contains a table of fifth roots and powers.) If $l$ is quite large, we may put $d = 0$ for a first approximation. In connection with these examples, see last figure. LONG PIPES. 713 EXAMPLE 1.—What head $h$ is necessary to deliver 190 cuin. ft. of water per minute through a clean straight iron pipe 140 ft. long and 6 in. in diameter? From eq. (2), with $v$, $b$, $h$, and sec., we have $$v = v_m = \left(4 \times \frac{V}{A} + \pi h\right) = 10.18 \text{ ft. per sec.}$$ Now for $s = 10$ ft. per sec. and $d = \frac{4}{3}$ ft., we find (in table, § 517) $f = .00549$; and hence, from eq. (3), $$h = \frac{(10.18)^2}{2 \times 32.2} \left[1 + .505 + \frac{4 \times .00549 \times 140}{\frac{4}{3}}\right] = 12.23 \text{ ft.,}$$ of which total head, as we may call it, 1.60 ft. is used in pro- ducing the velocity 10.18 ft. per sec. (i.e., $v_m + 2g$ = 1.60 ft.). while .088 ft. ($= c_x \frac{v_m^2}{2g}$) is lost at the entrance $E$ (with $\alpha$ = 90°), and .938 ft. (friction-head) is lost in skin-friction. EXAMPLE 2.—[Data from Weisbach.] Required the de- livery, $Q$, through a straight clean iron pipe 48 ft. long and 2 in. in diameter, with 5 ft. head (= $h$). $v$, $v_m$, being un- known, we first take $f = .006$ and obtain (eq. (4)) $$v_m = \sqrt{\frac{1}{1 + .505 + \frac{4 \times .00688 \times 48}{\frac{4}{3}}}} \cdot \frac{\sqrt{2} \times 32.2 \times 5}{\frac{4}{3}}$$ $$= 6.18 \text{ ft. per sec.}$$ From the table, § 517, for $v = 6.2$ ft. per sec. and $d = 2$ in., $$f = .00688,$$ whence $$v_m = \sqrt{\frac{1}{1 + .505 + \frac{4 \times .00688 \times 48}{\frac{4}{3}}}} \cdot \frac{\sqrt{2} \times 32.2 \times 5}{\frac{4}{3}}$$ $$= 6.04 \text{ ft. per sec.,}$$ which is sufficiently close. Then, for the volume per second, $$Q = \frac{\pi}{4} d^2 v_m = \pi (\frac{4}{3})^2 (.064) = 0.1367 \text{ cub. ft. per sec.}$$ 714 MECHANICS OF ENGINEERING. [Weibach's results in this example are $$v_m = 0.52 \text{ ft. per sec.}$$ $$Q = 0.1490 \text{ cub. ft. per sec.},$$ and his values for $f$ are slightly different.] **Example 8.** [Data from Weibach.] What must be the diameter of a straight clean iron pipe 100 ft. in length, which is to deliver $Q = \frac{1}{4}$ of a cubic foot of water per second under 5 ft. head (= A)? With $f = .006$ (approximately), we have from eq. (7), putting $d = 0$ under the radical for a first trial ($t$, lb., sec.), $$d = \sqrt{\frac{4 \times .006 \times 100}{2 \times 32.2 \times 5}} = \frac{1}{\pi} = \text{about } 0.30 \text{ ft.;}$$ whence $$v = \frac{4Q}{w d^3} = 7 \text{ ft. per sec.}$$ For $d = 3.8$ in. and $v = 7$ ft. per sec., we find $f = .00601$; whence, again, $$d = \sqrt{\frac{1.505 \times .30 + 4 \times .00601 \times 100}{2 \times 32.2 \times 5}} = \frac{4 \times .3}{\pi} = .3924 \text{ ft.;}$$ and the corresponding $v = 6.06$ ft. For this $d$ and $v$ we find $f = .00609$, whence, finally, $$d = \sqrt{\frac{1.505 \times .30 + 4 \times .00609 \times 100}{2 \times 32.2 \times 5}} = .398 \text{ ft.}$$ [Weibach's result is $d = .318$ ft.] **516. Chezy's Formula—If, in the problem of the preceding paragraph, the pipe is so long, and therefore $l : d$ so great, that $\frac{A}{l} : d$ in eq. (3) is very large compared with $1 + c_a$, we may neglect the latter term without appreciable error; whence eq. (3) reduces to $$h = 4f\left(\frac{l}{d}\right)\frac{v_m^*}{2g}.$$ (pipe very long; Fig. 579). . . . (8) A diagram showing a pipe with a head of water at one end and a flow rate at the other end. CHIEV'S FORMULA. 715 which is known as Chiev's Formula. For example, if $l = 1000$ ft. and $d = 2$ in. = 4 ft., and $f \approx 0.06$, we have $\frac{4}{d} \cdot \frac{l}{d} = 144$, while $1 + \xi$ for square corners is 1.508 only. If in (8) we substitute $$v_m = \frac{Q}{\sqrt{\pi d^3}} = Q + \frac{1}{2}\pi d,$$ (8) reduces to $$h = \frac{64}{\pi^2} \cdot \frac{l}{d^2} \cdot \frac{Q^2}{B_0} \cdots \quad (\text{very long pipe}); \quad \ldots \quad (9)$$ so that for a very long pipe, considering $f$ as approximately constant, we may say that to deliver a volume $Q$ per unit of time through a pipe of given length $l$, the necessary head, $h$, is inversely proportional to the fifth power of the diameter. And again, solving (9) for $Q$, we find that the volume conveyed per unit of time is directly proportional to the fifth power of the square root of the diameter; directly proportional to the square root of the head; and inversely proportional to the square root of the length. (Not true for short pipe; see above example.) If we conceive of the insertion of a great number of pheometers along the long straight pipe, of uniform section, now under consideration, the summits of the respective water columns maintained in them will lie in a straight line joining the discharging (into the air) end of the pipe with a point in the reservoir surface vertically over the inlet extremity (practically so), and the "slope" of this line (called the Hydrostatic Grade Line or Gradient), i.e., the tangent (or sine; the angle is so small, generally) of the angle which it makes with the horizontal is $\frac{1}{T}$, and may be denoted by $s$. Putting also $\frac{1}{d} R = r$ the hydraulic radius of the section of the pipe, and $v_m = v$ velocity in pipe, we may transform eq. (9) into $$v = \sqrt{\frac{2g}{f}} (Re)^{\frac{1}{2}}; \quad \text{or}, \quad v = A (Re)^{\frac{1}{2}}, \quad \ldots \quad (10)$$ 716 MECHANICS OF ENGINEERING. which is the form by which Mr. Hamilton Smith (see § 506) interprets all the experiments quoted by him on long pipes. As to notation, however, he uses $n$ for $A$, and $r$ for $R$. With the foot and second as units, the quantity $A$ (not an abstract number) varies approximately between 60 and 140. For a given $A$ we easily find the corresponding $f$ from the relation $$f = \frac{2g}{A}.$$ If the pipe discharges under water, $h$ = the difference of elevation of the two reservoirs. If the pipe is not horizontal, the use of the length of its horizontal projection instead of its actual length in the relation $s = \frac{2g}{A}$ occasions an error, but it is in most cases insignificant. Similarly, if a steady flow is going on in a long pipe of uniform section, at the extremities of any portion of which we have measured the piezometer heights (or computed them from the readings of steam or pressure gauges), we may apply eq. (9), putting for $h$ the difference of level of the piezometer summits, and for $l$ the length of the pipe between them. § 500. Coefficient $f$ in Fire-engine Hose.—Mr. Geo. A. Ellis, C.E., in his little book on "Fire-streams," describing experiments made in Springfield, Mass., gives a graphic comparison (p. 45) of the friction losses occurring in leather and rubber hose, in leather and iron pipe in clean iron pipe, each of 24 in. diameter, with various velocities; on which the following statements may be based: That for the given size of hose and pipe ($d = 24$ in.), the coefficient $f$ for the leather and rubber hose respectively may be obtained approximately by adding to $f$ for clean iron pipe (and a given velocity) the per cent of itself shown in the accompanying table.
Velocity per sec. Length per sec. Length per sec.
Velocity per sec. Length per sec. Length per sec.
$8.0$ $50\%$ $80\%$ $16$ $48\%$ $32\%$
$6.4$ $30\%$ $60\%$ $16$ $48\%$ $32\%$
$10$ $20\%$ $40\%$ $16$ $48\%$ $32\%$
$16$ $10\%$ $20\%$ $16$ $48\%$ $32\%$
**Example.**—For a clean iron pipe 24 in. in diameter for a velocity of 10 ft. per sec., we have, from § 517, $f =$ .00593. Hence for a leather hose of the same diameter, we have, for $v =$ 10 ft. per sec., $$f = .00593 + .00593 = .0088.$$ *For the most recent and exhaustive experiments in this direction see Mr. J. H. Freeman's "Hydraulics of Fire-engines" in the Transact Am. Soc. C. E. R. for 1889. (Also p. 582 of this work.)* A table showing coefficients of friction for different velocities and lengths of pipes. **PRESSURE-ENERGY.** 717 581. Bernoulli's Theorem as an Expression of the Conservation of Energy for the Liquid Particles.--In any kind of flow without friction, steady or not, in rigid immovable vessels, the aggregate potential and kinetic energy of the whole mass of liquid concerned is necessarily a constant quantity (see §§ 148 and 149), but individual particles (as the particles in the sinking free surface of water in a vessel which is rapidly being emptied) may be continually losing potential energy, i.e., reaching lower and lower levels, without any compensating increase of kinetic energy or of any other kind; but in a steady flow under ordinary conditions, we have to admit that the stock of energy of a given particle, or small collection of particles, is constant during the flow, provided we recognize a third kind of energy which may be called Pressure-energy, or capacity for doing work by virtue of internal fluid pressure; as may be thus explained: In Fig. 574 let water, with a very slow motion and under a pressure $p$ (due to the reservoir-head + atmosphere-head behind it), be admitted behind a piston into a space at rest in vacuo. Let $x =$ length of stroke, and $F =$ area of piston. At the end of the stroke, by motion of proper valves, communication with the reservoir is cut off on the left of the piston and opened on the right, while the water in the cylinder now on the left of the piston is put in communication with the vacuum exhausted by the return stroke. The pressure of this water falls to zero (height of cylinder small), and on the return stroke is simply conveyed out of the cylinder, neither helping nor hindering the motion. That is, in doing the work of one stroke, viz., $$W = \text{force} \times \text{distance} = Fp \times x = Fps,$$ a volume of water $V = Fs$, weighing $Fg$ (i.e., or other unit'), has been used, and, in passing through the motor, has experienced no appreciable change in velocity (motion slow), and 718 MECHANICS OF ENGINEERING. therefore no change in kinetic energy, nor any change of level, and hence no change in potential energy, but it has given up all its pressure. (See § 409 for y.) Now $W$, the work obtained by the consumption of a weight $G = V_y$ of water, may be written $$W = Fp_e = Fp_p = V_p = V_y \frac{p}{y} = G \frac{p}{y} \quad . \quad . \quad (1)$$ Hence a weight of water $G$ is capable of doing the work $G \times \frac{p}{y} = G \times$ head due to pressure $p$, i.e., $= G \times$ pressure-head, in giving up all its pressure $p$; or otherwise, while still having a pressure $p$, a weight $G$ of water possesses an amount of energy, which we may call pressure-energy, of an amount $$= G \frac{p}{y},$$ where $y$ is the heaviness (\(\S\) 7) of water, and $\frac{p}{y}$ is height, or head, measuring the pressure $p$; i.e., it equals the pressure-head. We may now state Bernoulli's Theorem without friction in a new form, as follows: Multiplying each term of eq. (1), § 451, by $Q_y$, the weight of water flowing per second (or other time-unit) in the steady flow, we have $$Q_y \frac{v^2_a}{2g} + Q_y \frac{p_a}{y} + Q_y z_a = Q_y \frac{v^2_b}{2g} + Q_y \frac{p_b}{y} + Q_y z_b \quad (2)$$ But $Q_y \frac{v^2_a}{2g} = \frac{1}{2} Q_y v^2_a = \frac{1}{2} \times mass\ flowing\ per\ time-unit \times square\ of\ the\ velocity = the\ kinetic\ energy\ inherent\ in\ the volume Q of water on passing the section m, due to the velocity at m. Also, $Q_y \frac{p_a}{y} = the\ pressure-energy\ of\ the\ volume Q$ at $m$, due to the pressure at $m$; while $Q_y z_a = the\ potential energy\ of\ the\ volume Q$ at $m$ due to its height $z_a$ above the arbitrary datum plane. Corresponding statements may be made for the terms on the right-hand side of (2) referring to the other section, $n$, of the pipe. Hence (2) may be thus read: *The aggregate amount of energy* (of the three kinds mentioned) *residient in the particles* of liquid when passing section n is **LOSS OF ENERGY.** 719 equal to that when passing any other section, as $n$ ; in steady flow without friction in rigid motionless vessels; that is, the store of energy is constant. 532. Bernoulli's Theorem with Friction, from the Standpoint of Energy.—Multiply each term in the equation of § 516 by $Qy$, as before, and denote a loss of head or height of resistance due to any cause by $A_r$, and we have $$Qy \frac{w^*}{2g} + Qy \frac{P_0}{v} + Qy h_r = Qy \frac{w^*}{2g} + Qy \frac{P_0}{v} + Qy h_r - Z_p Qy h_r \dots (3)$$ Each term $Qy h_r$ (e.g., $Qy \frac{w^*}{d} \frac{v^*}{d} 2g$) due to skin-friction in a long pipe, and $Qy \frac{w^*}{2g}$ due to loss of head at the reservoir entrance of a pipe represents a *loss of energy*, occurring between any position $n$ and any other position $m$ downstream from $n$, but is only in existence in the form of heat generated by the friction of the liquid particles against each other or the sides of the pipes. As illustrative of several points in this connection, consider two short lengths of pipe in Fig. 575, $A$ and $B$, one offering a gradual, the other a sudden enlargement of section, but otherwise identical in dimensions. We see that these occupy places in separate lines of pipe in each of which a steady flow with full cross-sections is proceeding, and so regulated that the velocity and internal pressure at $n$, in $A$, are equal respectively to those at $n$ in $B$. Hence, if vacuum pleometers be inserted at $n$, the A diagram showing two sections of pipe, A and B, with different cross-sectional areas. Section A has a gradual change in diameter, while section B has a sudden enlargement. The diagram also shows the velocity and pressure changes across these sections. Fig. 575. 720 MECHANICS OF ENGINEERING. smaller section, the water columns maintained in them by the internal pressure will be of the same height, $\frac{P_{n}}{y}$, for both $A$ and $B$. Since at $m$, the larger section, the sectional area is the same for both $A$ and $B$, since $F_{n}$ in $A = F_{n}$ in $B$, so that $Q_{A} = Q_{B}$, hence $v_{n}$ in $A = v_{n}$ in $B$ and is less than $v_{m}$. Now in $B$ a loss of head occurs (and hence a loss of energy) between $n$ and $m$, but now in $A$ (except slight friction-head); hence in $A$ we should find as much energy present at $m$ as at $n$, only differently distributed among the three kinds, while at $m$ in $B$ the aggregate energy is greater than that at $n$ in $A$. As regards kinetic energy, there has been a loss between $n$ and $m$, and in $A$ and $B$ (equal losses), but less than $v_{n}$. As to potential energy, there is no change between $n$ and $m$ either in $A$ or $B$, since at $n$ and $m$ are on a level. Hence if the loss of kinetic energy in $B$ is not compensated for by an equal gain of pressure-energy (as it is in $A$), the pressure-head $\left(\frac{P_{m}}{y}\right)$ at $m$ in $B$ should be less than that $\left(\frac{P_{n}}{y}\right)$ at $n$ in $A$. Experiment shows this to be true, the loss of head being due to the internal friction in the eddies occasioned by the sudden enlargement of section at $m$. At $m$, in $B$, there is a column of a less height than that at $m$ in $A$, whereas at $n$ they are equal. (See p. 467 of article "Hydromechanics" in the Ency. Bri. tannica for Mr. Frude's experiments). In brief, in A the loss of kinetic energy has been made up in pressure-energy, with no change of potential energy, but in B there is an actual absolute loss of energy = $\zeta Qh$, or $\zeta Qv\frac{v}{g}h$, suffered by the weight $Qv$ of liquid. The value of $\zeta$ in this case and others will be considered in subsequent paragraphs. Similarly, losses of head, and therefore losses of energy, occur at elbows, sharp bends, and obstructions, causing eddies and internal friction, the amount of each loss for a given weight, $G$, of water being = $\zeta G\frac{v^{2}}{g}h$, i.e., $\zeta = \frac{\zeta G\frac{v^{2}}{g}h}{G}$ being the loss of head occasioned by the obstruction (§ 474). It is SUDDEN ENLARGEMENTS IN PIPES. 721 therefore very important in transmitting water through pipes for purposes of power to use all possible means of preventing disturbance and eddying among the liquid particles. E.g., sharp corners, turns, elbows, abrupt changes of section, should be avoided in the design of the supply-pipe. The amount of the losses of head, or heights of resistance, due to these various causes will now be considered (except skin-friction, already treated). Each such loss of head will be written in the form $$\zeta = \frac{v^2}{2g}$$ and we are principally concerned with the value of the abstract number $\zeta$, or coefficient of resistance, in each case. The velocity $v$ is the velocity, known or unknown, where the resistance occurs; or if the section of pipe changes at this place, then $v =$ velocity on the down-stream section. The late Professor Weisbach, of the mining-school of Freiberg, Saxony, was one of the most noted experimenters in this respect, and will be frequently quoted. 533. Loss of Head Due to Sudden (i.e., Square-edged) Enlargement. Berna's Formula.—Fig. 576. An eddy is formed in the angle with consequent loss of energy. Since each particle of water of weight $G_s$, arriving with the velocity $v$, in the small pipe, may be considered to have an impact against the base of the infinitely great and more slowly moving column of water in the large pipe, and after the impact, move on with the same specific impulse that it had before, so that no energy is lost by elastic direct central impact ($\zeta$ 60), we may find the energy lost by this particle on account of the impact by eq. (1) of § 138, in which putting $M_s = G_s + g$, and $M_t = G_t + g =$ mass of infinitely great body of water in the large pipe, so that $M_s = w$, we have Energy lost by particle = $G_s \left( \frac{v - v_s}{2g} \right)^2$ and the corresponding Loss of head = $\frac{(v - v_s)^2}{2g}$ 729 MECHANICS OF ENGINEERING. which, since $F_v = F_{v_0}$, may be written $$\text{Loss of head in sudden enlargement} = \left(\frac{F}{F_0} - 1\right)^{\zeta} \cdot \text{ (9)}$$ That is, the coefficient $\zeta$ for a sudden enlargement is $$\zeta = \left(\frac{F}{F_0} - 1\right)^{\zeta} \cdot \text{ (3)}$$ $F_v$ and $F_0$ are the respective sectional areas of the pipes. Eq. (3) is Borda's Formula. Norm.—Practically, the flow cannot always be maintained with full sections. In any case, if we assume the pipes to be running full (once started so), and on that assumption compute the internal pressure at $F_v$, and find it to be zero or negative, the assumption is incorrect. That is, unless there is some pressure at $F_v$, the water will not swell out laterally to fill the large pipe. Example.—Fig. 577. In the short tube $AB$ containing a sudden enlargement, we have given $F_v = F_0 = 6$ sq. inches, $F_v = 4$ sq. inches, and $h = 9$ feet. Required the velocity of the jet at $m$ (in the air, so that $p_m + y = b = 34$ ft.), if the only loss of head considered is that due to the sudden enlargement (kinetic friction neglected, as the tube is short); the reservoir entrance has rounded corners. As shown in Fig. 577, we return to $m$ as down-stream section, and $n$ in reservoir surface as up-stream position (datum level at $m$), we have $$\frac{v_n^2}{2g} + b + 0 = 0 + b + h - \zeta \frac{v_n^2}{2g} \cdot \text{ (4)}$$ But, here, $v_n = v_m$ $$\therefore (1 + \zeta) \frac{v_n^2}{2g} = h \cdot \text{ (5)}$$ From eq. (3) we have $$\zeta = (\xi - 1)^{\zeta} = 0.25,$$ Fig. 577. SUDDEN ENLARGEMENT IN PIPE. 723 and finally (ft., lb., sec.) $$v_m = \sqrt{\frac{1}{1.25} \times 9 \times 32.2 \times 9} = 0.895 \sqrt{9 \times 32.2 \times 9}$$ $$= 21.55 \text{ ft. per sec.}$$ (The factor $0.895$ might be called a coefficient of velocity for this case.) Hence the volume of flow per second is $$Q = F_v v_m = r_f v_m \times 21.55 = 0.898 \text{ cub. ft. per sec.}$$ We have so far assumed that the water fills both parts of the tube, i.e., that the pressure $p_1$, at $F_1$, is greater than zero (see foregoing note). To verify this assumption, we compute $p_1$ by applying Bernoulli's Theorem to the centre of $F_1$ as downstream position and datum plane, and $w$ as upstream position, with no loss of head between, and obtain $$\frac{v_m^2}{2g} + p_1 + g = b + h - 0.$$ (6) But since $F_{y0} = F_v w$, we have $$v_m^* = (\frac{1}{y}) v_m^* = (\frac{1}{y}) w_m,$$ and hence the pressure-head at $F_1$ (substituting from equations above) is $$\frac{p_1}{y} = b + h - (\frac{1}{y})^2 \frac{h}{1 + c} = 84 + \frac{9}{4} + \frac{9}{4 + 25} = 97 \text{ feet},$$ and $\therefore p_1 = \frac{H}{y}$ of $14.7$ lbs. per sq. inch, which is greater than zero; hence efflux with the tube full in both parts can be maintained under $9$ ft. head. If, with $F_1'$ and $F_2'$ as before (and $\therefore C$), we put $p_1' = 0$, and solve for $h$, we obtain $h = 42.5$ ft. as the maximum head under which efflux with the large portion full can be secured. 524. Short Pipe, Square-edged Internally.—This case, already 794 MECHANICS OF ENGINEERING. treated in §§ 507 and 515 (see Fig. 578; a repetition of § 509), presents a loss of head due to the sudden enlargement from the contracted section at $m'$ (whose sectional area is $A$) to the full section $F$ of the pipe. From § 515 we know that the loss of head due to the short pipe is $h_s = \frac{C}{g} \left( \frac{a - m'}{m'} \right)^2$, in which $\frac{C}{g} = 0.568$ while from Borda's Formula, § 523, we have also $C = \left( \frac{F}{CF} - 1 \right)^2$. Equating these, we find the coefficient of internal contraction at $m'$ to be $$C = \frac{1}{1 + V_{\text{eff}}} = \frac{1}{1 + V_{\text{eff}}^{0.584}} = 0.584,$$ or about 0.60 (compare with $C = 0.64$ for thin-plate contraction, § 495). It is probably somewhat larger than this (§ 584), since a small part of the loss of head, $h_s$, is due to friction at the corners, and not all of this is due to contraction. By a method similar to that pursued in the example of § 523, we may show that unless $h_s$ is less than 40 feet, about, the tube cannot be kept full, the discharge being as in Fig. 551. If the efflux takes place into a "partial vacuum," this limiting value of $h_s$ is still smaller. Weibach's experiments confirm these statements (but those in the C. U. Hyd. Lab. seem to indicate that the limiting value for $h_s$ in the first case is about 50 ft.). 523. Diaphragm in a Cylindrical Pipe—Fig. 579. The diaphragm being of "thin plate," let the circular opening in it (concentric with the pipe) have an area $F_1$, while the sectional area of pipe is $F'$. Beyond $F'$, the stream contracts to a section of area $= CF' = F_1$, as shown by A diagram showing a cylindrical pipe with a circular opening in its wall, labeled \( F_1 \), and a stream contracting beyond \( F' \) to a section of area equal to \( CF' = F_1 \). Fig. 579 Let \( h_1 \) be the loss of head due to contraction at \( F_1 \), and \( h_2 \) that due to contraction at \( F' \). Then $$h_1 = \frac{C}{g} \left( \frac{a - m'}{m'} \right)^2,$$ and $$h_2 = \frac{C}{g} \left( \frac{a - m''}{m''} \right)^2,$$ where \( m'' \) is some intermediate point between \( m' \) and \( m'' \). Since \( h_1 + h_2 = h_s \), we have $$\frac{C}{g} \left( \frac{a - m'}{m'} \right)^2 + \frac{C}{g} \left( \frac{a - m''}{m''} \right)^2 = h_s.$$ **SHORT PIPE. DIAPHRAGM IN PIPE.** From which to fill the section $F_1$ of pipe, a loss of head occurs which by Borda's Formula, § 583, is $$h_s = \zeta \frac{v^2}{2g} = \left(\frac{F_1}{F_2} - 1\right) \frac{v^2}{2g}$$ where $v$, is the velocity of the pipe (supposed full). Of course $F_1$ (or $OF_1$) depends on $F_2$; but since experiments are necessary in any event, it is just as well to give the values of $\zeta$ itself, as determined by Weisbach's experiments, viz.
$\frac{F_1}{F_2}$ .10 .30 .40 .50 .60 .70 .80 .90 1.00
$\zeta$ = .298. 48. 17.5 7.8 3.7 1.5 4.0 3.8 3.5 3.8 3.6 3.4 3.6 3.4 3.2 3.4 3.2 3.0 3.2 3.0 2.8 3.0 2.8 2.6 2.8 2.6 2.4 2.6 2.4 2.2
By internal lateral filling, Fig. 580, the change of section may be made gradual and eddying prevented; and then but little loss of head (and therefore little loss of energy) occurs, besides the slight amount due to skin-friction along this small surface. On p. 467 of the article *Hydromechanics* in the *Encyclopaedia Britannica* may be found an account of experiments by Mr. Froods, illustrating this fact. **583. "The Venturi Water-meter."—The invention bearing this name was made by Mr. Clemens Herschel (see *Trans. Am Soc Civ Engineers*, for November 1887), and may be described as a method of pipe in which a gradual narrowing of section is accompanied successively by a more gradual enlargement, as in Fig. 580; but the diameter remains constant during the flow the piezometer-heights are observed at the three positions $r$, $n$, and $m$ (see below), and the rate of discharge may be computed as follows: Referring to Fig. 580, let us denote by $r$ (up-stream) position where the narrowing of the pipe begins, and by $n$ that where the enlargement ends, while $m$ refers to the narrowest section $F_m = F_r$. Applying Bernoulli's Theorem to sections $r$ and $n$, assuming 736 MECHANICS OF ENGINEERING. no loss of head between, we have, as the principle of the apparatus, $$\frac{p_1}{y} + \frac{v_1^2}{2g} = \frac{p_2}{y} + \frac{v_2^2}{2g}; \quad \ldots \quad (1)$$ whence, since $F_{e_0} = F_e e_0$, $$e_0 = \sqrt{\frac{1}{1 - (\frac{F_e}{F_0})^2}} \sqrt{2g(\frac{p_1}{y} - \frac{p_2}{y})} = \phi \sqrt{2g(\frac{p_1}{y} - \frac{p_2}{y})}, \quad (9)$$ in which $\phi$ represents the first radical factor. $\phi$ should differ but little from unity with $F_e$ small (and such was found to be the case by experiment). Its theoretical value is constant and greater than unity. In the actual use of the instrument the $E$ and $E_0$ are inferred from the observed piosmeter-heights $y$, and $y_0$ (since $\frac{p_1}{y} = y_c + b$, and $\frac{p_2}{y} = y_a + b$, $b$ being 34 ft.), and then the quantity flowing per time-unit computed, from $Q = F_e v_e$, $v_e$ having been obtained from eq. (9). This process gives a value of $Q$ about four per cent in excess of the truth, according to the second set of experiments mentioned below, if $v_e = 35$ ft. per sec.; but only one per cent excess with $v_e = 5$ or 6 ft. per sec. Experiments were made by Mr. Herschel on two meters of this kind, in each of which $F_e$ was only one ninth of $F_0$, a ratio so extreme that the loss of head due to passage through the instrument is considerable. E.g., with the smaller apparatus, in which the diameter at $a$ was 4 in., the loss of head between $r$ and $m$ was 10 or 11 ft., when the velocity through $a$ was 80 ft. per sec., those at other velocities being roughly proportional to the square of the velocity. In the larger instrument $d_a$ was 3 ft., and the loss of head between $r$ and $m$ was much the same proportion of its velocity as in the smaller (Eq. with $v_a = 34.56$ ft. per sec. the loss of head was 2.07 ft., while with $v_a = 19.96$ ft. per sec. it SUDDEN DIMINUTION OF SECTION OF PIPE. 737 was 0.49 ft.) The angle of divergence was much smaller in these meters than that in Fig. 580. 527. Sudden Diminution of Cross section, Square Edges.—Fig. 581. Here, again, the resistance is due to the sudden enlargement from the contracted section to the full section $F_1$ of the small pipe, so that in the loss of head, by Borda's formula, $$h_s = \zeta \frac{v^2}{2g} = \left( \frac{F_1}{F_2} - 1 \right) \frac{v^2}{2g} \dots (1)$$ the coefficient $$\zeta = \left( \frac{F_1}{F_2} - 1 \right) = \left( \frac{F_1}{CF_2} - 1 \right) = \left( \frac{1}{C} - 1 \right) \dots (2)$$ depends on the coefficient of contraction $C$ but this latter is influenced by the ratio of $F_1$ to $F_2$, the sectional area of the larger pipe, $C$ being about .80 when $F_2$ is very large (i.e., when the small pipe issues directly from a large reservoir so that $F_1 : F_2$ practically = 0). For other values of this ratio Weibach gives the following table for $C$, from his own experiments:
$For F_1 : F_2 = .10$ .30 .80 .40 .50 .60 .70 .80 .90 1.00
$C = .84$ .68 .68 .69 .69 .719 .755 .813 .899 1.00
$C$ being found, we compute $\zeta$ from eq. (3) for use in eq. (1). 528. Elbows.—The internal disturbance caused by an elbow, Fig. 582 (pipe full, both sides of elbow), occasions a loss of head $$h_s = \zeta \frac{v^2}{2g} \dots (1)$$ In which, according to Weibach's experiments with tubes 3 centims, i.e. 1.3 in., in diameter, we may put Fig. 583. 728 MECHANICS OF ENGINEERING.
For α = 20° 40° 60° 80° 90° 100° 110° 120° 130° 140°
ζ = 0.045 1.89 3.64 7.40 9.64 1.36 1.558 1.96 2.16 2.48
computed from the empirical formula; $$\zeta = 0.045 \sin^2 \frac{\alpha}{2} + 2.047 \sin^2 \frac{\alpha}{2}$$ σ is the velocity in pipe; σ as in figure. For larger pipes ζ would probably be somewhat smaller; and vice versa. If the elbow is immediately succeeded by another in the same plane and turning the same way, Fig. 583, the loss of head is not materially increased, since the eddying takes place chiefly in the further branch of the second elbow; but if it turns in the reverse direction, Fig. 584, but still in the same plane, the total loss of head is double that of one elbow; while if the plane of the second is $\frac{π}{4}$ to that of the first, the total loss of head is $\frac{π}{2}$ times that of one alone. (Weibach.) 589. Bends in Pipes of Circular Section.—Fig. 585. Weibach bases the following empirical formula for ζ, the coefficient of resistance, on his own experiments with pipes of circular section on his own experiments and some of Dubuat's, viz: $$\zeta = 0.131 + 1.847 \left( \frac{a}{r} \right)^2$$ (1) for use in $$h_v = v^2 \cdot \frac{a}{2g},$$ (9) where $a$ = radius of pipe, $r$ = radius of bend (to centre of pipe), and $v$ = velocity in pipe; $h_v$ = loss of head due to bend. ELBOWS AND BENDS IN PIPES. 720 It is understood that the portion $BC$ of the pipe is kept full by the flow; which, however, may not be practicable unless $BC$ is more than three or four times as long as wide, and is fall at the outlet. A semicircular bend occasions about the same loss of head as a quadrant bend, but two quadrants forming a re- verse curve in the pipe occasion a double loss. By enlarging the pipe at the bend, or providing internal thin partitions parallel to the sides, the loss of head may be considerably diminished. Weisbach gives the following table computed from eq. (1), but does not state the absolute size of the pipes.
$\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$ $\frac{a}{b}$
.10 .30 .60 .80 .100 .70 .80 1.00
Accounts of many of Weisbach's hydraulic experiments are contained in the *Civilingenieur*, vols. xx, x, and xl. 582a. Common Pipe elbows—Prof. L. F. Bellinger of Nor- wich University, Vermont, conducted a set of experiments in 1887, when a student at Cornell, on the loss of head occasioned by a common elbow (for wrought-iron pipe), whose longitudinal section is shown in Fig. 586a. The elbow served to connect at right angles two straight pipes having an internal diameter of .0485 in. A diagram showing a common pipe elbow. The internal diameter of the short bend or elbow was .4 in., and the radius of its curved circular axis (a quadrant) was .2 in. Its internal surface was that of an ordinary rough casting. A straight pipe of the same character and size and 14 feet long was first used, and the loss of head due to skin-friction (the only frictional loss in that case) carefully determined for a range of velocities from 2 to 90 ft./per second. Two lengths of similar pipe were then joined by the elbow 730 MECHANICS OF ENGINEERING. mentioned, forming a total length of 14 feet, and the total loss of head again determined through the same range of velocities. By subtraction, the loss of head due to the elbow was then easily found for each velocity, and assuming the form $$A = \zeta \frac{v^2}{2g} \quad \quad (1)$$ for the loss of head, $\zeta$ was computed in each case. From Fig. 586a it is seen that the stream meets with a sudden enlargement and a sudden diminution, of section, as well as with the short bend; so that the disturbance is of a rather complex nature. The principal results of Prof. Bellinger's experiments, after the adjustment of the observed quantities by "least squares," were found capable of being represented fairly well by the formula $$\zeta = 0.891 + [(\frac{v}{n})^2 - 1] \times 0.0876 \quad \quad (2)$$ where $n = [\text{veloc. in pipe in ft. per sec.}] + 5$. The following table was computed from eq. (2) (where $v$ is in ft. per second):
$n$ 9 4 6 8 10 12 14 16 18 20
$\zeta$ 0.883 0.69 0.70 0.67 0.74 0.75 0.85 0.99 1.089 1.185
530. Valve-gates and Throttle-valves in Cylindrical Pipes.— Adopting, as usual, the form $$A_v = \zeta \frac{v^2}{2g} \quad \quad (1)$$ for the loss of head due to a valve-gate, Fig. 587, or for a throttle-valve, Fig. 588, each in a definite position, Weibach's Fig. 587. Fig. 588. VALVE-GATE. THROTTLE-VALVE. 731 experiments furnish us with a range of values of $\zeta$ in the case of these obstructions in a cylindrical pipe 1.6 inches in diameter, as follows (for meaning of $x$, $d$, and $\alpha$, see figures. $v$ is the velocity in the full section of pipe, running full on both sides.)
Valve-gate. Throttle-valve.
$\frac{d}{d}$ $\zeta$
1.0 .00 .84
.07 10° .82
.26 90° 1.54
.81 90° 2.41
.81 90° 3.91
.06 45° .33
.53 90° .36
.53 90° .58
17.00 90° 11.90
.97.8 70° 731.
531. Examples involving Diverse Losses of Head.--We here suppose, as before, that the pipes are full during the flow. Practically, provision must be made for the escape of the air which collects at the high points. If this air is at a tension greater than one atmosphere, automatic air-valves will serve to provide for its escape; if less than one atmosphere, an air-pump can be used, as in the case of a siphon used at the Kansas City Water Works. (See p. 346 of the Engineering News for November 1887.) A diagram showing a valve gate and throttle valve in a pipe. Example 1.--Fig. 589. What head, $h$, will be required to deliver $\frac{1}{4}$ U.S. gallon (i.e. 231 cubic inches) per second 732 MECHANICS OF ENGINEERING. through the continuous line of pipe in the figure, containing two sizes of cylindrical pipe ($d_1 = 3$ in., and $d_2 = 1$ in.), and two 90° elbows in the larger. The flow is into the air at $m$, the jet being 1 in. in diameter like the pipe. At $E_1 = m$ 90°, and the constant velocity is called $v_1$. The common rounded. Use the ft.-lb.-sec. system of units in which $g = 32.2$. Since $Q = \frac{1}{8} gal = \frac{1}{8} \times 6068 = 0.068$ cub. ft. per sec., and therefore the velocity of the jet $$v_n = v_1 = Q + \frac{1}{2} m (v_1^2) = 12.95 ft. per sec.;$$ hence the velocity in the large pipe is to be $v_n = (\frac{1}{4}) v_1 = 1.36$ ft. per sec. From Bernoulli's Theorem, we have, with $m$ as datum plane, $$\frac{v_n^2}{2g} + b + 0 = 0 + b + h - C_x \frac{v_1^2}{2g} - 4f \frac{l_1}{d_2} \frac{v_1^2}{2g}$$ $$- 2C_a \frac{v_1^2}{2g} - C_x \frac{v_1^2}{2g} - 4f \frac{l_1}{d_2} \frac{v_n^2}{2g},$$ involving six separate losses of head, for each of which there is no difficulty in finding the proper $c$ or $f$, since the velocities and dimensions are all known, by consulting preceding paragraphs. (Clean iron pipe.) From § 515, table, for $\alpha = 90°$ we have . . . . $C_x = 0.505$ $\therefore$ $S_{517}$, for $d_1 = 3$ in., and $v_n = 13.88 ft. per sec., f$_x$ = .00728 $\therefore$ $d_2 = 1$ in., and $v_n = 13.25 ft.$ per sec., f$_x$ = .00613 $\therefore$ S58 (elbows), for $\alpha = 90°$ . . . . $C_d = 0.984$ $\therefore$ S587, for sudden diminution at $K$ we have [since $F_x + F_y = 1^* + b^* = 0.11$, : : $C_x = 0.695$] $$C_x = \left(\frac{1}{\sqrt{6}} - 1\right)^* = 0.360.$$ Solving the above equation for A, then, and substituting above numerical values (in ft.-lb.-sec.-system), we have (noting that $v_n = v_e$, and $v_e = v_1$) $$A = \frac{(19.25)^*}{64.4}\left(1 + (\frac{1}{4})(\frac{505 + 4 \times .00728 \times 50}{4} + 9 \times .984)\right.$$ $$+ .360 + \frac{4 \times .00613 \times .984}{\sqrt{6}}).$$ EXAMPLES; WITH LOSSES OF HEAD. I.e, $$A = \left(12.25^2\right) \left[1 + (0.0083 + 0.07160 + 0.0348) + (0.360 + 5.8848)\right];$$ $$A = 3.828 \times 7.3469 = 17.09 ft.$$ Ans. It is here noticeable how small are the losses of head in the large pipe, the principal reason of this being that the velocity in it is so small ($v_0$ = only 1.36 ft per sec), and that in general losses of head depend on the square of the velocity (nearly). In other words, the large pipe approximates to being a reservoir in itself. With no resistances a head $h = v_0^* + 2g$ ft would be sufficient. Example 2.—Fig. 590. With the valve-gate $V$ half raised (i.e., $s = \frac{1}{2}$ in Fig. 587), required the volume delivered per second through the clean pipe here shown. The jet issues A diagram showing a vertical pipe with a valve gate V at the top, connected to a horizontal pipe with a nozzle at the bottom. The nozzle has two branches, one labeled "q" and the other labeled "q". The diagram also shows a scale indicating distances. from a short straight pipe, or nozzle (of diameter $d_1 = \frac{1}{4}$ in.) inserted in the end of the larger pipe, with the inner corners not rounded. Dimensions as in figure. Radius of each bend $r = 2$ in. The velocity $v_a$ of the jet in the air is velocity $v_e$ in the small pipe; hence the loss of head at $K$ is $$\zeta_{x} \frac{v_a^*}{2g} = \zeta_x \frac{v_e^*}{2g}.$$ Now $a_e$ is unknown, as yet; but $v_e$, the velocity in the large pipe, is $= v_a \left(\frac{1}{2}\right)$; i.e., $v_e = \frac{1}{2} v_a$. From Bernoulli's Theorem: 739 734 MECHANICS OF ENGINEERING. orem (as datum level) we obtain, after transposition, $$h = \frac{v}{g} + C_2 \frac{v^2}{g} + C_3 \frac{v^2}{g} + 2C_4 \frac{v^2}{g} + 4f \frac{v^2}{g} + C_5 \frac{v^2}{g}$$ (1) Of the coefficients concerned, $f_c$ alone depends on the unknown velocity $v_c$. For the present [first approximation], put $$f_c = .006$$ From § 515, with $\alpha = 90^\circ$, $$C_2 = .505$$ From § 517, valve gate with $s = .0d$, $$C_3 = .206$$ From § 529, with $a : r = .5$, $$C_5 = .0294$$ While at $A_1$ from § 527, having $$(F_p/F_s)(F_p)^2 - F_s^2 = Y_s = .562;$$ we find from table, $$\ldots \ldots \ldots \ldots C = .700$$ and $$C_5 = \left(\frac{1}{700}\right)^2 = (0.428)^2$$ i.e., $C_5 = .0183$$ Substituting in eq. (1) above, with $v_c = (F_p/F_s)v_m$, we have $$v_m = \sqrt{\frac{1}{1 + \frac{81}{256} \left[ C_2 + 2C_3 + 4f_c \frac{v^2}{g} + C_5 \right]}} v_{ph},$$ (2) in which the first radical, an abstract number, might be called a coefficient of velocity, $\phi_v$ for the whole delivery pipe; and also, since in this case $Q = F_m v_m = F_n n$, may be written $Q = uF_n v_{ph}$, it may be named a coefficient of efflux, $\mu$. Hence $$v_m = \sqrt{\frac{1 + .421 v_{ph}}{.300 + .06 - .334 + (.300 - .06)}} v_{ph}$$ :. $v_m = .421 v_{ph} = .421 \times 39.9 \times .35 = 16.89$ ft. per sec. (The A3I might be called a coefficient of velocity.) The volume delivered per second is $$Q = \frac{1}{4} d^2 v_m = \frac{1}{4} \pi (F_p)^2 16.89 = .207$ cubic ft. per sec. (As the section of the jet $F_n = F_p$, that of the short pipe or nozzle, we might also say that $.421 = c_{eff}$ coefficient of efflux, for we may write $Q = uF_n v_{ph}$, whence $\mu = .421$.) FLOW THROUGH Siphons. 735 532. Siphons.—In Fig. 532, § 490, the portion $HN_1O$ is above the level $BC$, of the surface of the water in the head reservoir $BL_1$, and yet under proper conditions a steady flow can be maintained with all parts of the pipe full of water, including $HN_1O$. If the atmosphere exerted no pressure, this would be impossible; but its average tension of 14.7 lbs. per sq. inch is equivalent to an additional depth of nearly 34 feet of water placed upon $BC$. With no flow, or a very small velocity, the pipe may be kept full if $N_1$ not more than 33 or 34 feet above $BC$; but the greater $v_s$, the velocity of flow at $N_1$, the greater must be the height of water above head between $L$ and $N_1$, the less must be the height of $N_1$ above $BC$ for a steady flow. The analytical criterion as to whether a flow can be maintained or not, supposing the pipe completely filled at the outset, is that the internal pressure must be > 0 at all parts of the pipe. If on the superposition of a flow through a pipe of given design the pressure $p$ is found < 0, i.e. negative, at any point ($N_1$ being the important section for test) the supposition is inadmissible, and hence a steady flow cannot exist. For example, Fig. 532, suppose $LN_1N_2$ to be a long pipe of uniform section (diameter = $d$, and length = $l$), and that under the assumption of filled sections we have computed $v_s$, the velocity of the jet at $N_1$; i.e., $$v_s = \sqrt{\frac{1}{1 + \zeta_d + 4f\frac{l}{d}}} \quad . . . . . . . (1)$$ To test the supposition apply Bernoulli's Theorem to the surface $BC$ and the point $N_1$, where the pressure is $p_s$, velocity $v_s (= v_s)$, since we have supposed a uniform section for whole pipe), and height above $BC = A_s$. Also, let length of pipe $LN_1HN_2 = l_s$. Whence we have $$\frac{v_s^*}{2g} + p_s + h_s = 0 + \frac{P_n}{y} (b = b) + 0 - \zeta_d \frac{v_s^*}{2g} - 4f\frac{l}{d}\frac{v_s^*}{2g} \quad . . . . . . . (2)$$ [BC being datum plane.] A diagram showing a siphon system with various points labeled. 736 MECHANICS OF ENGINEERING. Solving for $p_1$, we have $$\frac{p_1}{r} = 34 \text{ feet} - \left[ A + \frac{v^*}{2g} (1 + C_2 + 4f) \right] \dots (5)$$ We note, then, that for $p_1$ to be > 0, $$h_m must be < \left[ 34 \text{ feet} - \frac{v^*}{2g} - C_2 - 4f \frac{l}{d g} \right]. \quad (4)$$ In the practical working of a siphon it is found that atmospheric air, previously dissolved in the water, gradually collects at $N_1$, the highest point, during the flow and finally, if not removed, will cause the water to rise. One device for removing the air consists in first allowing it to collect in a chamber in communication with the pipe beneath. This communication is closed by a stop-cock after the water in it has been completely displaced by air. Another stop-cock, above, being now opened, water is poured in to replace the air, which now escapes. Then the upper stop-cock is shut and the lower one opened. The same operation is again necessary until some air remains. On p. 346 of the Engineering News of November 1887 may be found an account of a siphon which has been employed since 1875 in connection with the water-works at Kansas City. It is 1350 ft. long, and transmits water from the river to the artificial "well," from which the pumping engines draw their supply. At the highest point, which is 16 ft. above low-water level of the river, is placed a "vacuum chamber" in which the air collects under a low tension corresponding to the height, and a pipe is connected with it to remove this air and prevent its breaking off of the (partly submerged) diameter of the pipe is 24 in., and the extremity in the "well" dips 5 ft. below the level of low water. See Troutwine's Pocket-book, for an account of Maj. Crozet's Siphon. 539a. Branching Pipes.—If the flow of water in a pipe is caused to divide and pass into two others having a common BRANCHING PIPES. 727 junction with the first, or vice versæ; or if lateral pipes lead out of a main pipe, the problem presented may be very complicated. As a comparatively simple instance, let us suppose that a pipe of diameter $d$ and length $l$ leads out of a reservoir, and at its extremity is joined to two others of diameters $d_1$ and $d_2$, and lengths $l_1$ and $l_2$, respectively, and that the further extremities of the latter discharge into the air without nozzles under heads $h_1$ and $h_2$ below the reservoir surface. Call these two pipes Nos. 1 and 2. Then the system forms a pipe in plan. Assuming that all entrances and junctions are smoothly rounded, so that all loss of head due to skin-friction, it is required to find the three velocities of flow, $v$, $v_1$, and $v_2$, in the respective pipes. First applying Bernoulli's Theorem to a stream-line from the reservoir surface through the main pipe to the jet at the discharging end of pipe No. 1, we have $$\frac{u^*}{2g} = h - 4f \frac{l}{d} \frac{v^*}{2g} - 4f \frac{l_1}{d_1} \frac{v_1^*}{2g}; \quad \ldots \quad (1)$$ and similarly, dealing with a stream-line through the main pipe and No. 2, $$\frac{u^*}{2g} = h_1 - 4f \frac{l}{d} \frac{v^*}{2g} - 4f \frac{l_2}{d_2} \frac{v_2^*}{2g}; \quad \ldots \quad (2)$$ while the equation of continuity for this case is $$\frac{du^*}{dt} = u d^* v + u d^* v_1 + u d^* v_2; \quad \ldots \quad (3)$$ From these three equations, assuming the same in all pipes as a first approximation, we can find the three velocities (best by numerical trial, perhaps); and then the volume of discharge of the system per unit of time $$Q = \frac{du^*}{dt} v; \quad \ldots \quad (4)$$ 533. Time of Emptying Vertical Prismatic Vessels (or Inclined Prisms of Bottom) by Variable Head. Case I. There is an orifice or short pipe in the bottom and opening into the air—Fig. 591. As the upper free surface 738 MECHANICS OF ENGINEERING. of area = $F'$, sinka, $F'$ remains constant. Let $z$ = head of water at any stage of the emptying; it = 0 at the outset, and = 0 when the vessel is empty. At any instant, $Q$, the rate of discharge (= volume per time-unit) depends on $z$ and is $$Q = \mu F \sqrt{2gz} \cdot \cdot \cdot (1)$$ where $\mu$ = coefficient of efflux = $\phi C$ - coefficient of velocity × coefficient of contraction [see § 495, eq. (8)]. We here suppose $F'$ so large compared with $F$, the area of the orifice, that the free surface of the water in the vessel does not acquire any notable velocity at any stage, and that hence the rate of efflux is the same at any instant, as for a steady flow with head $z$ = 0 and a new velocity in the free surface. $\mu$ is considered constant. From (1) we have $$dV = (\text{vol. discharged in time } dt) = Qdt = \mu F \sqrt{2gz} dt \cdot (2)$$ But this is also equal to the volume of the horizontal lamina $F'dz$, through which the free surface has sunk in the same time $dt$ $$\therefore - F'dz = \mu F \sqrt{2gz} dt; \quad dt = -\frac{F'}{\mu F \sqrt{2gz}} z - dz \cdot (3)$$ We have written minus $F'dz$ because, dt being an increment, $dz$ is a decrement. To reduce the depth from $z_s$ (at the start, time $t$ = zero) to $z_a$, demands a time $$\int_0^{t} dt = -\frac{F'}{\mu F \sqrt{2gz}} \int_{z_s}^{z_a} z - dz = \frac{2F'}{\mu F \sqrt{2gz}} [z_t - z] \cdot (4)$$ whence, by putting $z_s$ = 0, we have the time necessary to empty the whole prism $$\int_0^{t} dt = \frac{2F'z_t}{\mu F \sqrt{2gz}} = \frac{2F'z_a}{\mu F \sqrt{2gz}} = \frac{2 \times \text{volume of vessel}}{\text{initial rate of discharge}} \cdot (5)$$ TIME OF EMPTYING VESSELS. 730 that is, to empty the vessel requires double the time of discharging the same amount of water if the vessel had been kept full (at constant head $z = z_0$ = altitude of prism). To fill the same vessel through an orifice in the bottom, the flow through which is supplied from a body of water infinitely extensive, normally with the single normal line of Fig. 592, will obviously require the same time as given in eq. (5) above, since the effective head $z$ diminishes from $z_0$ to 0, according to the same law. **EXAMPLE.** What time will be needed to empty a parallelopipedal tank (Fig. 591) 8 ft. by 6 ft. in horizontal section and 6 ft. deep, through a stop-cock in the bottom, when the area of efflux with fully open is known to be $A = 0.64$, and whose rate of discharge is a circle of diameter = 3 in. t. From given dimensions $F^e = 4 \times 5 = 90$ sq. ft., while $z_0 = 6$ ft. Hence from eq. (5) (ft.-lb.-sec.) $$\text{time of emptying} = \frac{9 \times 20 \times \sqrt{6}}{0.64 \times 1^2(8t)} \times \frac{1}{32.2} = \frac{13980}{90} \text{ seconds}$$ Case II. Two communicating prismatic vessels. Required the time for the water to come to a common level $OX$, Fig. 593, efflux taking place through a small orifice, of area $F'$, under water. At any instant the rate of discharge is $$Q = mF' \sqrt{2gz},$$ as before. $z$ = difference of level. Now if $F'$ and $F''$ are the horizontal sectional areas of the two prismatic vessels (axes vertical) we have $F'' = F'^n$, and hence $x$, which is $x + y$, $= x + (F'^n + F'^n)y$ $$\therefore z = -\frac{s}{1 + F'^n}, \text{ and } ds = \frac{ds}{1 + F'^n}.$$ 740 MECHANICS OF ENGINEERING. As before, we have $$-F' \, dz = \mu F' \sqrt{2g} \, dh, \quad \text{or} \quad dt = -\frac{F'F''}{F' + F''} \mu F' \sqrt{2g} \, dh.$$ Hence, integrating, the time for the difference of level to change from $z_1$ to $z_2$ is $$= \frac{2F'F''}{F' + F''} \frac{z_2 - z_1}{\mu F' \sqrt{2g}}. \quad (6)$$ and by making $z_1 = 0$ in (6), we have the time of coming to a common level is $$= \frac{2F'F''}{F' + F''} \frac{1}{\mu F'} \sqrt{\frac{2g}{3g}}. \quad (7)$$ ALGEBRAIC EXAMPLE.—In the double lock in Fig. 594, let $L'$ be full, while in $L''$ the water stands at a level $MN$ the same as that of the tail-water. $F'$ and $F''$ are the horizontal sectional areas of the prismatic locks. Let the orifice, $O$, between them be at a depth $h$, below the initial level $KL$. Let $L''$ be filled to a height $h_1$, observe that, $MN$ of $L''$. The orifice at $O$, area = $F'_o$, being opened, efflux from $L'$ begins into the air, and the level of $L''$ is gradually raised from $MN$ to $OD$, while that of $L'$ sinks from $KB$ to $AB$ a distance = $a$, computed from the relation vol. $Fa = vol. F'' h_1$, and the time occupied is [eq. (4)] $$t_1 = \frac{2F'}{\mu F' \sqrt{2g}} (\sqrt{h_1} - \sqrt{h} - a). \quad (8)$$ As soon as $O$ is submerged, efflux takes place under water, and we have an instance of Case II. Hence the time of reaching a common level (after submergence of $O$) (see eq. 7) is $$t_2 = \frac{2F'F''}{\mu F'(F' + F'')} \sqrt{\frac{h_1 - a}{3g}}, \quad (9)$$ and the total time is = $t_1 + t_2$, with $a = F'' h_1 + F''$. TIME OF EMPTYING VESSELS. 741 CASE III. Emptying a vertical prismatic vessel through a rectangular "notch" in the side, or overfall.—Fig. 595. As before, let even the initial area (= $z b$) of the notch be small compared with the horizontal area $A'$ of tank. Let $s =$ depth of lower half of notch below level of tank surface at any instant, and $b =$ width of notch. Then, at any instant (see eq. 10, § 508), $$\text{Rate of disch. (vol.)} = Q = \frac{3}{2}\mu b \sqrt{2g s} = \frac{3}{2}\mu b \sqrt{2g} s$$ :. vol. of disch. in dt = $\frac{3}{2}\mu b \sqrt{2g} s dt$ and putting this is $-F' ds = \text{vol. of water lost by the tank in time } dt$, we have $$dt = -\frac{3}{2}\mu b \sqrt{2g} s^{-1} ds$$ whence $$\int dt = -\frac{3}{2}\mu b \sqrt{2g} \int s^{-1} ds = -\frac{3}{2}\mu b \sqrt{2g} \int_0^s s^{-1} ds$$ i.e., $$\int_0^t dt = -\frac{3F'}{\mu b \sqrt{2g}} \left[ \ln s + \frac{1}{2} \right]_0^s$$ (10) as the time in which the tank surface sinks from a height $z_a$ above sill to a height $z_a$ above sill. If we inquire the time $t'$ for the water to reach a height $z_a$ above sill, we put $s_a = z_a$, whence $t' = t$. As explanatory of this result, note that as $x$ diminishes not only does the velocity of flow diminish, but the available area of efflux ($= z b$) also grows less, whereas in Cases I and II the orifice of efflux remained of constant area = $F'$. Eq. (10) is applicable to the waste-weir of a large reservoir or pond. 534. Time of Emptying Vessels of Variable Horizontal Sections.—Considering regular geometrical forms first, let us take 742 MECHANICS OF ENGINEERING. CASE I. Wedge-shaped vessel, edge horizontal and under- neath, office $F'$ in the wedge that $z_0$, variable head is always the altitude of a triangle similar to the section $ABC$ of the body of water when efflux begins. At any instant during the efflux the area, $S_t$, of the free surface, variable here, takes the place of $F'$ in eq. (3) of § 533, whence we have, for any case of variable free surface, $dt = \frac{-Sa - dz}{\mu F' \sqrt{g}}$. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (11) In the present case $S = wu$, and from similar triangles whence $$S = b z + z_s,$$ and $$dt = \frac{-b z dz}{\mu F' \sqrt{g}}.$$ :. $$\int_{z_0}^{\infty} dt = \frac{b u}{\mu F' \sqrt{g}} \int_{z_0}^{\infty} dz = \frac{4b u}{\mu F' \sqrt{g}} \left[ z_t - z_s \right],$$ . . (19) and hence the time of emptying the whole wedge, putting $z_s = 0$, is $$t_c = \frac{4}{3} \frac{4b u}{\mu F' \sqrt{g}} = \frac{4}{3} \frac{\text{Vol. of wedge}}{\text{initial rate of discharge}},$$ . . (13) i.e., so long as to discharge the same volume of water under a constant head $= z_s$. This is equally true if the ends of the wedge are oblique, so long as they are parallel. CASE II. Right segment of paraboloid of revolution—Fig. 597. Axis vertical. Ori- fice at vertex. Here the variable free surface has at any instant an area, $S_t = wu$, $u$ being the radius of the circle and variable. From a property of the parabola $$w': b': s': z': : S = \pi b^2 + z_s,$$ Fig. 596. Fig. 597. TIME OF EMPTYING VESSELS. 763 and hence, from eq. (11), $$dt = - \frac{\pi b^2}{\mu F_0} zdz;$$ $$\therefore \int_{s_0}^{s_1} dt = \frac{\pi b^2}{\mu F_0} \int_{s_0}^{s_1} zdz = \frac{9}{8} \frac{\pi b^2}{\mu F_0} [z_1 - z_0];$$ whence, putting $s_a = 0$, we have the time of emptying the whole vessel $$t_a = \frac{4}{3} \frac{\pi b^2 F_0}{\mu F_0} = \frac{4}{3} \text{ initial rate of discharge}. \quad (14)$$ same result as for the wedge, in Case I; in fact, it applies to any vessel in which the areas of horizontal sections vary directly with their heights above the origin. Case III. Any pyramid or cone ; vertex down ; small office in vertex—Fig. 598. Let area of the base $= S_a$ at upper edge of vessel. At any stage of the flow $S = area$ of pyramid of water. From similar pyramids $$S_a : S :: s_a : s' ; \therefore S = \frac{S_a}{s'} s';$$ and [eq. (11)] $$dt = - \frac{S_a}{s'} \frac{1}{\mu F_0} zdz,$$ whence ($s_a = 0$) the time of emptying the whole vessel is $$t_a = \frac{S_a}{\mu F_0} \int_{s_0}^{s_1} zdz = \frac{9}{8} \frac{S_a}{\mu F_0};$$ or, $$t_a = \frac{6}{5} \text{ Total volume}. \quad (15)$$ Case IV. Sphere.—Similarly, we may show that to empty Fig. 598. 744 MECHANICS OF ENGINEERING. a sphere, of radius = $r$, through a small orifice, of area = $F$, in lowest part, the necessary time is $$t_4 = \frac{16}{15} \frac{\pi r^3}{\mu F \sqrt{gr}} = \frac{8}{5} \frac{\text{Vol.}}{\text{init. rate of disch.}}$$ 535. Time of Emptying a Obelisk-shaped Vessel.—(An obelisk may be defined as a solid of six plane faces, two of which are rectangles in parallel planes and with sides respectively parallel, the others trapezoids; a frustum of a pyramid is a particular case.) A volume of this shape is of common occurrence; see Fig. 599. Let the altitude = $h$, the two rectangular faces being horizontal, with dimensions as in figure. By traversing through $FG$, and $H$ right lines parallel to $EC$, to cut the upper base, we form a rectangle $KLMC$ equal to the lower base. Produce $ML$ to $P$ and $KL$ to $N$, and join $PG$ and $NG$. We have now subdivided the solid into a parallelepiped $PLMNHG$, a pyramid $PNBLG$, and two wedges, viz. $APLKH$ and $LNDMFG$, with their edges horizontal; and may obtain the time necessary to empty the whole obelisk-volume by adding the times which would be necessary to empty the individual component volumes, separately, through the same orifice or pipe in the bottom plane $EG$. These have been already determined in the preceding paragraphs. The dimensions of each component volume are given in terms of those of the obelisk, and all have a common altitude = $h$. Assuming the orifice to be in the bottom, or else that the discharging end of the pipe, if such is used, is in the plane of the bottom $EG$, we have as follows, $F$ being the area of discharge: Fig. 599 TIME OF EMPTYING RESERVOIRS. 745 Time to empty the parallelopiped $\{ \}$. $t_1 = \frac{2h}{\mu F \sqrt{2g}}$ (1) Time to empty the two wedges separately would be (Case I, § 533) $\{ \}$. $t_2 = \frac{2}{3} \frac{l_1(l_1 - l_2)(b_1 - b_2)}{\mu F \sqrt{2g}}$ (9) For the pyramid $\{ \}$... $t_3 = \frac{2}{5} \frac{(l - l_1)(b - b_1)}{\mu F \sqrt{2g}}$ (8) Hence to empty the whole reservoir we have $$t = t_1 + t_2 + t_3;$$ i.e., $$t = [3h + 8h_1 + 2h_2 + 2h_3] \cdot \frac{2\sqrt{k}}{15\mu F \sqrt{2g}}.$$ (4) EXAMPLE.—Let a reservoir of above form, and with $b = 50$ ft., $l = 60$ ft., $b_1 = 10$ ft., $l_1 = 20$ ft., and depth of water $h = 18$ ft., be emptied through a straight iron pipe, horizontal, and leaving the side of the reservoir close to the bottom, at an angle $\alpha = 36^\circ$ with the inner plane of side. The pipe is 80 ft. long and 4 inches in internal diameter; and of clean surface. The jet issuing directly from the pipe is 4 in. in diameter, and has a velocity of $v^2 = 4\pi^2 f^2$ sq. feet. To find "the coefficient of efflux" (= $\phi$, the coefficient of velocity in this case, since there is no contraction at discharge orifice), we use eq. (4) (the first radical) of § 518, with $f$ approx. = .006, and obtain $$\mu = \phi \sqrt{\frac{1}{1 + 4\pi^2 f^2/d}} = \sqrt{\frac{1 + .006 + 4\pi.006\times80}{d}} = .361.$$ (N.B. Since the velocity in the pipe diminishes from a value $$v = .361\sqrt{2g\times16} = 11.6$$ ft. per sec. at the beginning of the flow to zero at the close, $f$ = .006 is a reasonably approximate average with which to compute the average $\phi$ above; see § 517.) 740 MECHANICS OF ENGINEERING. Hence from eq. (4) of this paragraph (ft.-lb.-sec. system) $$t = \frac{3 \times 50 \times 60 + 8 \times 10 \times 90 + 2(50 \times 20 + 10 \times 60)}{15 \times 0.361} \sqrt{\frac{g}{\rho}} \sqrt{\frac{v^2}{g} \times 32.2}$$ = 29110 sec. = 8 hrs. 5 min. [ Probably within 2 or 3% of the truth.] 388. Time of Emptying Reservoirs of Irregular Shape. Simpson's Rule.--From eq. (1), § 384, we have, for the time in which the free surface of water in a vessel of any shape whatever sinks through a vertical distance $d_z$, $$dt = -\frac{S e^{-kz}}{\mu V^2 g z} dz,$$ whence $$\int_{z_0}^{z_1} dt = \frac{1}{\mu V^2 g z} \int_{z_0}^{z_1} S e^{-kz} dz,\quad . . . (1)$$ where $S$ is the variable area of the free surface at any instant, and $z$ the head of water at the same instant, efflux proceeding through a small orifice (or extremity of pipe) of area $F$. If $S$ can be expressed in terms of $z$, we can integrate eq. (1) (i.e., provided that $Se^{-kz}$ has a known antiderivative); but if not, the vessel or reservoir being irregular in form, as in Fig. 600 (which shows a pond whose bottom has been accurately surveyed, so that we know the value of $S$ for any stage of the emptying), we can still get an approximate solution by using Simpson's Rule for approximate integration. Accordingly, if we inquire the time in which the surface will sink from $z_0$ to the entrance $E$ of the pipe in Fig. 600 (any point $z$; at $E$ or short of that), we divide the vertical distance from $z_0$ to $z$ (in this figure) into an even number of equal parts, and from the known form of the pond compute the area $S$ corresponding to each point of division, giving them $\delta_1$, $\delta_2$, etc. Then the required time is approximately Fig. 600. 9/22 TIME OF EMPTYING POND. 747 $$t = \frac{x_0 - x_n}{8\mu F V_{2g} n} \left[ S_1 + 4 \left( \frac{S_1}{x_1^2} + \frac{S_2}{x_2^2} + \cdots + \frac{S_{n-1}}{x_{n-1}^2} \right) + 2 \left( \frac{S_1}{x_1^3} + \frac{S_2}{x_2^3} + \cdots + \frac{S_{n-1}}{x_{n-1}^3} \right) + \frac{S_n}{x_n^4} \right]$$ EXAMPLE.—Fig. 600. Suppose we have a pipe $Em$ of the same design as in the example of § 535, and an initial head of $z_0 = 16$ ft., so that the same value of $\mu = 361$, may be used. Let $x_0 - z_n = 8$ feet, and divide this interval (of $4$ ft.) into four equal parts, and let us find the area of each part by means of points of division we find from a previous survey that $S_0 = 400000$ sq. ft., $S_1 = 380000$ sq. ft., $S_2 = 270000$ sq. ft., $f_t = 210000$ sq. ft., and $S_n = 180000$ sq. ft.; while $a = 4$, $\mu = 361$, and the area $F = 4\pi(4)^3 = 0.873$ sq. ft., we obtain $(ft._n lb._n sec.)$ $$t = \frac{16 - 8}{0.361 \times 0.873 \times 2 \times 32.3 \times 3 \times 4} \left[ \frac{400000}{\sqrt{16}} + \frac{4 \times 380000}{\sqrt{16}} + \frac{9 \times 270000}{\sqrt{16}} + \frac{4 \times 210000}{\sqrt{16}} + \frac{180000}{\sqrt{8}} = 944400 sec. \right.$$ The volume discharged, $V$, may also be found by Simpson's Rule, thus: Since each infinitely small horizontal lamina has a volume $$dV = - Sdz, ~\therefore ~V = \int_{s_0}^{s_n} Sdz,$$ or, approximately, $$V = \frac{x_0 - x_n}{8n} \left[ S_1 + 4S_1 + 2S_2 + 4S_2 + \cdots + S_n \right].$$ Hence with $n = 4$ we have $(ft._n lb._n sec.)$ $$V = \frac{16 - 8}{8 \times 4} \left[ 40000 + 4 \left( \frac{38000}{+27000} + \frac{9 \times 27000}{+18000} + 2 \times 27000 } + 18000 = 2,160,000$ cub. ft. 748 MECHANICS OF ENGINEERING. 537. Volume of Irregular Reservoir Determined by Observing Progress of Emptying -- Transforming eq. (11), § 534, we have $$Sdt = - \mu F \sqrt{g} dt.$$ But $Sdt$ is the infinitely small volume $dV$ of water lost by the reservoir in the time $dt$, so that the volume of the reservoir between the initial and final ($0$ and $n$) positions of the horizontal free surface (at beginning and end of the time $t_n$) may be written $$\boxed{V = \mu F \sqrt{g} \int_0^{t_n} dt.}$$ (1) This can be integrated approximately by Simpson's Rule, if the whole time of emptying, $= t_n$, be divided into an even number of equal parts, and the values $x_1, x_2, x_3, ...$ of the ordinates noted at these equal intervals of time (not of vertical height). The corresponding surface planes will not be equidistant, in general. Whence for the particular case when $n = 4$ (see Fig. 601) $$\boxed{V = \frac{\mu F \sqrt{g}}{3 \times 4} (x_1 + 4(x_1 + x_2) + 2x_3 + x_4).}$$ (2) Diagram showing a water reservoir with a horizontal free surface. Fig. 601 CHAPTER VII. HYDRODYNAMICS (Continued)---STEADY FLOW OF WATER IN OPEN CHANNELS. 538. Nomenclature.--Fig. 602. When water flows in an open channel, as in rivers, canals, mill-races, water-courses, ditches, etc., the bed and banks being rigid, the upper surface is free to conform in shape to the dynamic conditions of each case, which therefore regulate to that extent the shape of the cross-section. In the vertical transverse section $AC$ in figure, the line $AC$ is called the air-profile (usually to be considered horizontal and straight), while the line $AB$, profile of the bed and banks, is called the wetted perimeter. It is evident that the ratio of the wetted perimeter to the whole perimeter, though never $\frac{1}{2}$, varies with the form of the transverse section. In a longitudinal section of the stream, $EFGH$, the angle made by a surface filament $EF$ with the horizontal is called the slope, and is measured by the ratio $i = h : l$, where $l$ is the length of a portion of the filament and $h$ is the fall, or vertical distance between two ends of a line. The length between the horizontal and the line $HG$ along the bottom is not necessarily equal to that of the surface, unless the portion of the stream forms a prism; i.e., the slope of the bed does not necessarily $= s = \text{that of surface}$. EXAMPLES.--The old Creton Aqueduct has a slope of 1.10 ft. per mile; i.e., $s = .00208$. The new aqueduct (for New A diagram showing a cross-sectional view of a river channel with a sloped bed and banks. 749 750 MECHANICS OF ENGINEERING. York) has a slope $s = .000138$, with a larger transverse section. For large sluggish rivers $s$ is much smaller. 539. Velocity Measurements.—Various instruments and methods may be employed for this object, some of which are the following: *Surface floats* are small balls, or pieces of wood, etc., so colored and weighted as to be readily seen, and still but little affected by the wind. These are allowed to float with the current in different parts of the width of the stream, and the surface velocity $v$ in each experiment computed from $c = l + t$, where $l$ is the distance described between parallel transverse alignments (or actual ropes where possible), whose distance apart is measured on the bank, and $t$ the time occupied. *Density floats.* Two balls of different densities, one lightest, the other heavier than water, are united by a slender chain, their weights being so adjusted that the light ball, without projecting notably above the surface, buoyes the other ball at any assigned depth. Fig. 603. It is assumed that the combination moves with a velocity $v'$ equal to the arithmetic mean of the surface velocity $v_s$ of the stream and that, $v'$ is the water filament at the depth of the lower ball, which latter, $v_s$ is generally less than $v'$. That is, we have $$v' = \frac{v_s + v'}{2} \quad \text{and} \quad v_s = \frac{2v' - v}{2} \dots (1)$$ Hence, $v_s$ having been previously obtained, eq. (1) gives the velocity $v'$ at any depth of the lower ball, $v'$ being observed. The *floating staff* is a hollow cylindrical rod, of adjustable length, weighted to float upright with the top just visible. Its observed velocity is assumed to be an average of the velocities of all the filaments lying between the ends of the rod. Woltmann's Mill; or Tachometer; or Current-meter. Fig. 604, consists of a small wheel with inclined float (or of a small A diagram showing a floating staff instrument. Fig. 603 CURRENT-METERS, ETC. 751 " screw-propeller " wheel S) held with its plane T to the current, which causes it to revolve at a speed nearly proportional to the velocity, \( v \), of the water passing it. By a screw-gearing W on the shaft, connection is made with a counting apparatus to record the number of revolutions. Sometimes a vanes B is attached, to compel the wheel to face the current. It is either held at the extremity of a pole or, by being adjustable along a vertical staff fixed in the bed, may be set at any desired depth below the surface. It is usually so designed as to be thrown in and out of gear by a cord and spring, that the time of making the indicated number of revolutions may be exactly noted. By experiments in currents of known velocities, formulae have been derived to interpret the indications of any one instrument; i.e., to find the velocity \( v \) of the current corresponding to an observed number of revolutions per minute. A peculiar form of this instrument has been recently invented, called the Ritchie-Haskell Direction-current Meter, for which the following is claimed: "This meter registers electrically on dial a boat from which used, the direction and velocity of any current. Can be used in river, harbor, or ocean currents." Pilot's Tube consists in principle of a vertical tube open above, while its lower end, also open, is bent horizontally upstream; see A in figure. After the oscillations have ceased, the water in the tube remains A diagram showing a "screw-propeller" wheel S held with its plane T to the current. The wheel is connected to a counting apparatus via a screw-gearing W on the shaft. A vanes B is attached to compel the wheel to face the current. Fig. 304. A B 752 MECHANICS OF ENGINEERING. stationary with its free surface a height $h'$, above that of the stream, on account of the continuous impact of the current against the lower end of the column. By the addition of another vertical tube (see B in figure) with the face of its lower (open) end parallel to the current (so that the water-level in it is the same as that of the current), both tubes being provided with stop-cocks, we may, after closing the stop-cocks, lift the apparatus into a boat and read off the height $h$ at level. We may also cause both columns of water to move through the tube by introducing air into both tubes in the boat by putting the upper ends of both tubes in communication with a receiver of rarefied air, and thus watch the oscillations and obtain a more accurate value of $h'$. [See Van Nostrand's Mag. for Mar. '78, p. 355.] Theoretically (see § 565), the thickness of the walls of the tube at the lower end being considerable, we have $$c = \sqrt{gh'} \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)$$ as a relation between $c$, the velocity of the particles impinging on the lower end, and the static height $A'$ (§ 565). Eq. (1) is verified fairly well by Weibach's experiments with fine instruments, used with velocities of from 0.32 to 1.24 meters per second. Weibach found $$c = \left[3.54 \sqrt{A'}\right] \text{ (in meters)} \quad \text{met. per sec.,}$$ whereas eq. (1) gives $$c = \left[3.138 \sqrt{A'}\right] \text{ (in meters)} \quad \text{met. per sec.}$$ In the instruments used by Weibach the end of the tube turning up-stream was probably straight; i.e., neither flaring nor conically convergent. A change in this respect alters the relation between $c$ and $A'$; see § 565 for Pitot's and Darcy's results. Pitot's Tube, though simple, is not so accurate as the u-ethometer. CURRENT-METERS. 753 The *Hydrometric Pendulum*, a rather uncertain instrument, is readily understood from Fig. 606. The side $AB$, of the quadrant $ABC$, being held vertical, the plane of the cord is inclined to the current. The angle $\theta$ between the cord and the vertical depends on $G$, the effective weight (i.e., actual weight diminished by the buoyant effort) of the ball (heavier than water), and the amount of $P$, the impulse or horizontal pressure of the current against the latter, since the cord will take the direction of the resultant $R$, for equilibrium. Now $P$ (see § 572) for a ball of given size and character of surface varies (nearly) as the square of the velocity; i.e., if $P'$ is the impulse on a given stationary ball, when the velocity of the current is $c'$, then for any other velocity $c$ we have $$P = \frac{P'}{c'^2} c^2 \quad \ldots \ldots \quad (3)$$ From this and the relation $\tan \theta = \frac{P}{Q}$ we derive $$c = \sqrt{\frac{Gc'}{P'}} \cdot \sqrt{\tan \theta} \quad \ldots \ldots \quad (3)$$ With a given instrument and a specified system of units, the numerical value of the first radical may be determined as a single quantity, by experimenting with a known velocity and the value of $\theta$ then indicated, and may then, as a constant factor, be employed in (3) for finding the value of $c$ for any observed value of $\theta$; but the same units must be used as before. 540. Velocities in Different Parts of a Transverse Section.—The results of velocity-measurements made by many experimenters do not agree in supporting any very definite relation between the greatest surface velocity ($c_{\text{max}}$) of a transverse A diagram showing a pendulum-like apparatus with a ball attached to a string, which is hanging over a horizontal rod. The ball is shown at rest, with its string forming an angle with the vertical. The label "Fig. 606" is below the diagram. Fig. 606 754 MECHANICS OF ENGINEERING. section and the velocities at other points of the section, but establish a few general propositions: 1st. In any vertical line the velocity is a maximum quite near the surface, and diminishes from that point both toward the bottom and toward the surface. 3d. In any transverse horizontal line the velocity is a maximum near the middle of the stream, diminishing toward the banks. 3d. The mean velocity = v, of the whole transverse section, i.e., the velocity which must be multiplied by the area, $F$, of the section, to obtain the volume delivered per unit of time, $$Q = Fv, \quad \ldots \quad \ldots \quad \ldots \quad (1)$$ is about 83 per cent of the maximum surface velocity ($v_{\text{max}}$) observed when the air is still; i.e., $$v = 0.83 \times (v_{\text{max}}). \quad \ldots \quad \ldots \quad (9)$$ Of eight experimenters cited by Prof. Bowser, only one gives a value (= 0.62) differing more than .05 from 83, while others obtained the values .82, .78, .82, .80, .82, .83. In the survey of the Mississippi River by Humphreys and Ablott, it was found that the variation of the velocity in any given vertical line could be fairly well represented by the ordinates of a parabola (Fig. 607) with its axis horizontal and its vertex at a distance $d$, below the surface according to the following relation, $f'$ being a number dependent on the force of the wind (from 0 for no wind to 10 for a hurricane): $$d = [0.317 + 0.06 f'] d; \quad (3)$$ where $d$ is the total depth, and the double sign is to be taken + for an upstream, - for a down-stream wind. The following relations were also based on the results of the survey: (putting, for brevity, $B = 1.69 + \frac{d}{d+1.5}$) $$\ldots \quad (4)$$ Fig. 607. **VELOCITIES IN OPEN CHANNEL.** $$e = c_e - \sqrt{Bw\left(\frac{z-d}{d}\right)^2}, \quad \ldots \quad \ldots \quad (5)$$ $$c_m = 4c_e + 4c_d + \frac{q}{d}(4c_e - 4c_d), \quad \ldots \quad (6)$$ and $$c_{\text{tot}} = c_m + \frac{v}{d}\sqrt{Bw}. \quad \ldots \quad (7)$$ (These equations are not of homogeneous form, but call for the foot and second as units.) In (4), (5), and (6), $$c = \text{velocity at any depth } z \text{ below the surface};$$ $$c_m = \text{mean velocity in the vertical curve};$$ $$c_d = \max. \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \quad " \\ c_{\text{mid}} = "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" "" \\ c_b = "\text{velocity at bottom}"; \\ v = "\text{mean velocity of the whole transverse section}." It was also found that the parameter of the parabola varied inversely as the square root of the mean velocity $c_m$ of curve. In general the bottom velocity ($c_b$) is somewhat more than $\frac{3}{4}$ the maximum velocity ($c_m$) in the same vertical. In the Misissippi the velocity at mid-depth in any vertical is found to be very nearly half of the surface velocity in the same vertical; this fact is important, as it simplifies the approximate gauging of a stream. 541. Gauging a stream or river.—Where the relation (eq. (9), § 540) $v = 83(c_m - c_d)$ is not considered accurate enough for substitution in $Q = Fv$ to obtain the volume of discharge (or delivery) $Q$ of a stream per time-unit, the transverse section may be divided into a number of subdivisions as in Fig. 608, of widths $a_1$, $a_2$, etc., and mean depths $d_1$, $d_2$, etc., and respective mean velocities, $c_1$, $c_2$, etc., computed from measurements with current-meters; whence we may write $$Q = a_1d_1c_1 + a_2d_2c_2 + a_3d_3c_3 + etc..$$ (Fig. 608) A diagram showing a cross-section of a stream with various widths and depths labeled \(a_1, a_2, a_3\) and corresponding depths \(d_1, d_2, d_3\) respectively. (7) 756 MECHANICS OF ENGINEERING. With a small stream or ditch, however, we may erect a vertical boarding, and allow the water to flow through a rectangular notch or overfall, Fig. 609, and after the surface has become permanent, measure $a$, (depth of sill below the level surface somewhat back of boards), and $b$ (width) and use the formula of § 504; see examples in that article. 542. Uniform Motion in an Open Channel.—We shall now consider a straight stream of indefinite length in which the flow is steady, i.e., a state of permanency exists, as distinguished from a freshet or a wave. That is, the flow is steady when the water assumes fixed values of mean velocity $v$, and sectional area $F$, on passing a given point of the bed or bank; and the Eq. of continuity . . . $Q = Fv = Fv = Fv = constant$ . . . (1) holds good whether those sections are equal or not. By uniform motion is meant that (the section of the bed and banks being of constant size and shape) the slope of the bed, the quantity of water (volume of $Q$) flowing per time-unit, and the extent of the wetted perimeter, are so adjusted to each other that the mean velocity of flow is the same in all transverse sections, and consequently the area and shape of the transverse section is the same at all points; and the slope of the surface = that of the bed. We may therefore consider, for simplicity, that we have to deal with a prism of water of indefinite length sliding down a plane surface of constant slope and moving with uniform velocity (via, i.e., mean velocity $v$ common to all the sections); that is, there is no acceleration. Let Fig. 610 show, free, a portion of this prism, of length $l$, and having its bases $\gamma$ to the bed and surface. A diagram showing a cross-section of an open channel with a straight line indicating uniform flow. Fig. 609. UNIFORM MOTION. OPEN CHANNEL. 757 The hydrostatic pressures at the two ends balance each other from the identity of conditions. The only other forces having Diagram showing a prism with water on one side and air on the other. PERFECT PERFECT Fig. 82. components parallel to the bed and surface are the weight $G = F \cdot y$ of the prism (where $y$ = heaving of water) making an angle $\epsilon$ (= slope) with a normal to the surface, and the friction between the water and the bed which is parallel to the surface. The amount of this friction for the prism in question may be expressed as in § 510, viz: $$P = fric. = f \cdot S \cdot y \cdot \frac{v^2}{2g} = f \cdot poly \cdot \frac{v^2}{2g}, \quad \ldots \quad (9)$$ in which $S = w \cdot l$ = rubbing surface (area) = wetted perimeter, $w$, × length (see § 538), and $f$ an abstract number. Since the mass of water in Fig. 610 is supposed to be in relative equilibrium, we may apply it to the laws of motion of a rigid body, and since the motion is a uniform translation (§ 109) the components, parallel to the surface, of all the forces must balance. :. $G \sin \epsilon$ must $= P = fric.; \therefore F \cdot y = f \cdot poly \cdot \frac{v^2}{2g};$ whence $$h = f \cdot poly \cdot \frac{v^2}{F^2} \cdot \frac{l}{2g}, \quad \ldots \ldots \ldots (8)'$$ or $$h = f' \cdot R \cdot \frac{v^2}{2g}, \quad \ldots \ldots \ldots (3)$$ in which $F + w$ is called $R$, the hydraulic mean depth, or hydraulic radius. (3) is sometimes expressed by saying that 768 MECHANICS OF ENGINEERING. the whole fall, or head, $h$, is (in uniform motion) absorbed in friction-head. Also, since the slope $s = h + l$, we have $$v = \sqrt{\frac{2g}{f}} \cdot \sqrt{hl}; \text{ or, } v = A \cdot \sqrt{hl}, \dots (4)$$ which is of the same form as Chebyshev's formula in § 519 for a very long straight pipe (the slope $s$ of the actual surface in this case corresponding to the slope along pionometer-summits that of a closed pipe). In (4) the coefficient $A = \sqrt{\frac{2g}{f}}$ is not, like $f$, an abstract number, but its numerical value depends on the system of units employed. 542a. Experiments on the flow of water in open channels—Those of Darby and Bazin, begun in 1855 and published in 1865 ("Recherches Hydrauliques"), were very carefully conducted with open conduits of a variety of shapes, sizes, slope, and character of surface. In most of these a uniform flow was secured before the taking of measurements. The velocities ranged between from about 0.5 to 8 or 10 ft. per second; the hydraulic radii from 0.03 to 3.0 ft., with deliveries as high as 182 cub. ft. per second. For example, the following results were obtained in the canals of Marseilles and Cropano; the quantity $Q$ being for the foot and second. The sections were nearly all rectangular. See eq. (4) above.






















































































<
                                                                           &nb... UNIFORM MOTION. KUTTER'S FORMULA. 750 good mortar joints, and about 9 ft. wide; the depths of water ranging from 1.5 to 4.5 ft. With plaster of pure cement on the bed in one of the experiments the high value of $A = 133.6$ was reached (foot and second), with $v = 808$ ft. per second, $R = 2111$ ft. per second, and $s = 87.1$ cm. ft. per second. Captain Cunningham, in his experiments on the Ganges Canal at Roorkee, India, in 1881, found $A$ to range from 48 to 130 (foot and second). Humphreys and Abbot's experiments on the Mississippi River and branches (see § 540), with values of $R =$ from 2 or 3 ft. to 72 ft., furnish values of $A =$ from 53 to 167 (foot and second). § 549. Kutter's Formula.—The experiments upon which Weibisch based his deductions for $f$, the coefficient of fluid friction, were scanty and on too small a scale to warrant general conclusions. That author considered that $f$ depended only on the velocity, disregarding altogether the degree of roughness of the bed, and gave a table of values in accordance with this view, varying from $f = .0002$ at 2 ft. per sec. to $.0109$ for 0.4 ft. per sec.; but in 1869 Moser, Kutter and Ganguillet, having a much wider range of experimental data at command, including those of Darcy and Bazin, and those obtained on the Mississippi River, evolved a formula, known as Kutter's Formula, for the uniform motion of water in open channels, which is claimed to harmonize in a fairly satisfactory manner the chief results of the best experiments in that direction. They make the coefficient $A$ in eq. (4) (or rather the factor $\sqrt{\frac{v}{g}}$ contained in $A$) a function of $R$, $s$, and also $n$ an abstract number, or coefficient of roughness, depending on the nature of the surface of the bed and banks ; viz., $$\begin{array}{rlr} & {v \text{ in } \\ & \text{ft.} \\ & \text{per sec.}} & {= \left[ \frac{41.6 + \frac{1.811}{n} + \frac{0.0092}{s}}{1 + \left( \frac{41.6 + \frac{0.0092}{s}}{n} \right) \sqrt{\frac{s}{R(\text{in ft.)}}} } \right] \sqrt{\frac{s}{R(\text{in feet})}} } & {(b)} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {} \\ & {} & {}\end{array}$$ which is Kutter's Formula. 780 MECHANICS OF ENGINEERING. That is, comparing (5) with (4), we have $f$ a function of $n$, $R$, and $s$, as follows: $$f = \frac{1 + 41.6 + 0.0021}{\sqrt{R} \text{ in ft.}} \cdot \cdot \cdot \cdot \cdot (6)$$ From (6) it appears that $f$ decreases with an increasing $R$, as has been also noted in the case of closed pipes (§ 517); that it increases with increasing roughness of surface; and that it is somewhat dependent on the slope. The makers of the formula give the following values for $n$. Values of $n$. $n =$ .009 for well-planned timber bed; .010 for plaster in pure cement; .011 for plaster in cement with $\frac{3}{4}$ sand; .012 for unplanted timber; .013 for brickwork;* .015 for canvas lining on frames; .017 for rubble; .020 for canal in very firm gravel; .025 for rivers and canals in perfect order and regimen, and perfectly free from stones and weeds; .030 for rivers and canals in moderately good order and regimen, but overgrown with vegetation; .035 for rivers and canals in bad order and regimen, overgrown with vegetation and strewn with stones or detritus of any sort. Kutter's Formula is claimed to apply to all kinds and sizes of watercourses, from large rivers to sewers and ditches; for uniform motion. If $\sqrt{V R}$ is the unknown quantity, Kutter's Formula leads to a quadratic equation; if $s$ the slope, to a cubic. Hence, to save computation, tables have been prepared, some of which will be found in vol. 28 of Van Nostrand's Magazine. * For ordinary brick sewers Mr. B. F. Hartford claims that $n = .04$ gives good results. See Jour. Eng. Societies for '84-'85, p. 290.
No. (cub. ft. per sec.) $R$ (ft.) Area numb. 0% per sec.) (foot and sec.) Character of the
1 183.78 1.004 0.0087 10.36 137.1 Very smooth.
2 183.78 1.004 0.0087 10.36 137.1 Very smooth.
3 48.98 7.06 0.099 11.38 79.4 Rough.
4 48.98 7.06 0.099 11.38 79.4 Rough.
5 48.98 891 -0.0212 7.98 78.5 Rather rough.
6 48.98 -0.0212 -0.0212 -7.98
$f$ $= \frac{1 + 41.6 + 0.0021}{\sqrt{R} \text{ in ft.}}$
$n$ Values
UNIFORM MOTION. OPEN CHANNEL. 761 (pp. 135 and 398) (sewers), and in Jackson's works on Hydraulics (rivers). The following table will give the student an idea of the variation* of the coefficient $A = \sqrt{\frac{g}{R}}$, of eq. (4), or large bracket of eq. (5), with different hydraulic radii, slopes, and values of $n$, according to Kutter's Formula; from $R = 4$ ft., for a small ditch or sluice-way (or a wide and shallow stream), to $R = 15$ ft., for a river or canal of considerable size. Under each value of $R$ are given two values of $A$; one for a slope of $s' = .001$, and the other for $s' = .00005$. All these values of $A$ imply the use of the foot and second. These values of $A$ have been scaled by the writer from a diagram given in Jackson's translation of Kutter's "Hydraulic Tables," and are therefore only approximate. The corresponding value of $\nu$, the coefficient of fluid friction, can be computed from $\nu = \frac{59}{A^2}$.
n $R = 4$ ft. $R = 1$ ft. $R = 8$ ft. $R = 16$ ft.
0.000 133 148 187 174 187 196 199 222
0.001 133 148 187 174 187 196 199 222
0.002 58 69 70 69 70 69 70 69
0.003 58 69 70 69 70 69 70 69
0.004 31 43 43 40 58 58 66 72
0.005 31 43 43 40 58 58 66 72
The formula used in designing the New Aqueduct for New York City, in 1855, by Mr. Fieley, consulting engineer, was [see (4)] $$v (\text{ft. per sec.)} = 149 \sqrt{R (\text{in ft.)} \times s} \quad \ldots \quad (7)$$ whereas Kutter's Formula gives for the same case (a circular section of 14 ft. diameter, and slope of 0.7 ft. to the mile), with $n = 0.013$, $$v (\text{ft. per sec.)} = 140.7 \sqrt{R (\text{in ft.)} \times s} \quad \ldots \quad (8)$$ *The aqueduct has this circular form for a small portion, only, of its length; a "hose-like" section of very nearly the same flowing capacity being given to the greater portion of the remainder. A table showing values of A and R for different n values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showing values of v for different R and s values. A table showingvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRandsvaluesofvfordifferentRstandscalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescalescale 782 MECHANICS OF ENGINEERING. To quote from a letter of Mr. I. A. Shaler of the Aqueduct Corps of Engineers, "Mr. Freley states that the cleanliness of the conduit (Sudbury) had much to do in affecting the flow. He found the flow to be increased by 7 or 8 per cent in a portion which had been washed with a thin wash of Portland cement." EXAMPLE 1.—A canal 1000 ft. long of the trapezoidal section in Fig. 611 is required to deliver 300 cubic ft. of water per second with the water 8 ft. deep at all sections (i.e., with uniform motion), the slope of the bank being such that for a depth of 8 ft. the width of the water surface (or length of air-profile) will be 20 ft.; and the coefficient for rounding off being $n = .020$. What is the necessary slope to be given to the bed (slope of bed = that of surface, here) (ft./ft., sec.)? The mean velocity $$v = Q + F = 300 + \frac{1}{2}(20 + 8) \cdot 8 = 2.67 \text{ ft. per sec.}$$ (So that the surface velocity of mid-channel in any section would probably be $(c_{\text{max}}) = v + 0.83 = 3.51 \text{ ft. per sec. (eq. (3), § 540).}$) The wetted perimeter $$w = 8 + 2 \sqrt{8^2 + 6^2} = 28 \text{ ft.,}$$ and therefore the mean hydraulic depth $$= R + w = 112 + 28 = 4 \text{ ft.}$$ To obtain a first approximation for the slope, we may use the value $f = .00795$ given by Weisbach for a velocity of 2.67 ft. per sec., and obtain, from (3), $$h = \frac{.00795 \times 1000 \times 98(2.67)^2}{112 \times 3 \times 32.2} = .291 \text{ ft.;}$$ i.e., $s = h + l = .00291.$ UNIFORM MOTION IN OPEN CHANNEL 763 With this value for the slope and $R = 4$ ft. (see above), we then have, from eq. (6) (putting $n = .020$), $$f = \frac{1 + (41.6 + .0081 + .090)}{5.184 + .020 + .000221} = .0071,$$ with which value of $f$ we now obtain $$A = 0.300 \text{ feet}; \quad \text{i.e., slope } m = .0090.$$ EXAMPLE 2.—If the bed of a creek falls 20 inches every 1500 ft. of length, what volume of water must be flowing to maintain a uniform mean depth of 4 ft., the corresponding surface-width being 40 ft., and wetted perimeter 46 ft.? The bed is "in moderately good order and regimen;" use Kutter's Formula for finding $n = .089$ (ft. and sec.). First we have $$\sqrt{Rt} = \sqrt{(40 \times 4)} \div \left(46 \times \frac{1500}{4}\right) = .066,$$ while $\sqrt{R}(ft.) = 1.98$, and the slope $s = \frac{H}{L} + 1500 = .00111$; hence $$v = \frac{41.6 + .089 + .091}{1 + \frac{.0981}{.00111}} \times .066 = \frac{104.43 \times .066}{1.6685},$$ or $$v = 4.18 \text{ ft. per sec.}$$ Hence, also, $$Q = Fv = 40 \times 46 \times 4.18 = 743.4 \text{ cub. ft. per sec.}$$ [N.B. Weibach works this same example by eq. (8) with a value of $f$ taken from his own table, his result being $v = 6.1$] 764 MECHANICS OF ENGINEERING. ft. per sec., which would probably be attained in practice only by making the bed and tanks smoother than as given.] EXAMPLE 8.—The desired transverse water-velocity in a canal is to be 3 ft. in 1600 ft.; i.e., $v = \frac{3}{1600}$; or, for $l = 1600$ ft., $h = 3$ ft. What must be the velocity (mean) of each section, for a uniform motion ; the corresponding volume delivered per sec., $Q = F_v = t$; assuming that the character of the surface warrants the value $n = .030$ ? Knowing the slope $s$, $= \frac{3}{1600}$; and the hydraulic radius $R$, $= F + w$, or $75.80$ sq. ft. $= 34.97$ ft., $= 5.215$ feet; with $n = .030$ we substitute directly in eq. (8), obtaining $v = 4.97$ ft. per sec.; whence $Q = F_v = 370$ cub. ft. per sec. 543. Hydraulic Mean Depth for a Minimum Frictional Resistance.—We note, from eq. (8), § 549, that if an open channel of given length $l$ and sectional area $F$ is to deliver a given volume, $Q$, per time-unit with uniform motion, so that the common mean velocity of all sections ($= Q/F$) is also a given quantity, the necessary fall $h$, or slope $s = h/l$, is seen to be inversely proportional to $R$, the hydraulic mean depth of the section, $(F + w)$, sectional area + wetted perimeter. For $h$ to be as small as possible, we may design the form of transverse section, so as to make $R$ large as possible; i.e., to make the wetted perimeter a minimum for a given $F$; for in this way a minimum of frictional contact, or area of rubbing surface, is obtained for a prism of water of given sectional area $F$ and given length $l$. In a closed pipe running full the wetted perimeter is the whole perimeter; but if the given sectional area is shaped in the form of a circle, the wetted perimeter, i.e., $w$, is a minimum (and $R$ a maximum). If the full pipe must have a polygonal shape of $n$ sides, then the regular polygon of $n$ sides will provide a minimum $w$. Wherein it follows that if the pipe or channel is running UNIFORM MOTION IN OPEN CHANNEL. 785 half full, and thus becomes an open channel, the semicircle, of all curvilinear water profiles, gives a minimum $w$; but all rectangular sections with sides at 60° with the horizontal half of a regular hexagon gives a minimum $w$. Among all rectangular sections the half square gives a minimum $w$; and of all half octagons the half of a regular octagon gives a minimum $w$ (and max. $R$) for a given $F$. See Fig. 618 for all these. The egg-shaped outline, Fig. 614, small end down, is frequently given to sewers in which it is important that the different velocities of the water at differ- ferent stages (depths) of flow (depend- ing on the volume of liquid passing per unit-time) should not vary widely from each other. The lower portion $ABC$, providing for the lowest stage of flow, $AB$, is nearly semicircular, and thus in- duces a velocity at its lowest stage (being constant at all stages) which does not differ extremely from that occurring when the water flows at its highest stage $DE$, although this latter velocity is the greater; the reason being that $ABC$ from its advantageous form has a hydraulic radius, $R$, larger in proportion to its sectional area, $F$, than $DCE$. That is, $F \cdot w$ for $ABC$ is more nearly equal to $F \cdot w'$ for $DEC$ than if $DEC$ were a semicircle, and the velocity at the lowest stage may still be sufficiently great to prevent the de- posit of sediment. See § 575. 544. Trapezoid of fixed side-slope.—For large artificial water- courses and canals the trapezoid, or three-sided water-profile (symmetrical), is customary, and the inclination of the bank, A diagram showing a trapezoidal cross-section of an open channel. Fig. 613 A diagram showing a semicircular cross-section of an open channel. Fig. 614 766 MECHANICS OF ENGINEERING. or angle $\theta$ with the horizontal, Fig. 615, is often determined by the nature of the material composing it, to guard against washouts, caving in, etc. We may thus be concerned with the following problem. Given the area, $F$, of the transverse section, and the angle $\theta$, required the value of the depth $x$ (or of upper width $z$, or of lower width $y$, both of which are functions of $z$) to make the hydraulic mean depth, $R = F + w$, a maximum, or so $Fw$ a minimum. $F$ is constant. From the figure we have $$w = AB + 2BF = y + 2z\sec \theta, \quad \ldots \quad (1)$$ and $$F = yz + x^2\cot \theta; \quad \ldots \quad (2)$$ whence $$y = \frac{1}{z}(F - x^2\cot \theta). \quad \ldots \quad (3)$$ substituting which in (1) and dividing by $F$, noting that $2\sec \theta - \cot \theta = \frac{2-\cos \theta}{\sin \theta}$, we have $$\frac{w}{F} = \frac{1}{z} + \frac{2-\cos \theta}{F\sin \theta}. z. \quad \ldots \quad (3)$$ For a minimum $w$ we put $$\frac{d(w)}{dx} = 0; \text{ i.e., } -\frac{1}{xz} + \frac{2-\cos \theta}{F\sin \theta} = 0;$$ $$\therefore x(\text{for max. or min. } w) = \pm\sqrt{\frac{F\sin \theta}{2-\cos \theta}}.$$ The $+$ sign makes the second derivative positive, and hence for a min. $w$ or max. $R$ we have $$x(\text{call it } x') = x' = -\sqrt{\frac{F\sin \theta}{2-\cos \theta}}. \quad (4)$$ TRAPEZOID FOR MINIMUM WETTED PERIMETER. 767 while the corresponding values for the other dimensions are $$y' = \frac{F}{x'} - x'\cot{\theta} \quad \ldots \quad \ldots \quad \ldots \quad (5)$$ and $$x' = y' + 2x'\cot{\theta} = \frac{F}{x'} + x'\cot{\theta} \quad \ldots \quad \ldots \quad (6)$$ For the corresponding hydraulic mean depth $R'$ [see (3)], i.e., the max. $R$, we have $$\frac{1}{R'} = \frac{1}{x'} + \frac{2 - \cos{\theta}}{F\sin{\theta}} = \frac{2}{x'}; \quad \ldots \quad (7)$$ $$\therefore R' = 4x' = \frac{1}{2}\sqrt{\frac{F\sin{\theta}}{2 - \cos{\theta}}} \quad \ldots \quad (8)$$ Equations (4), (5), $\ldots$ (8) hold good, then, for the trapezoidal section of least frictional resistance for a given angle $\theta$. Problem.-Required the dimensions of the trapezoidal section of minimum frictional resistance for $\theta = 45^\circ$, which with $A = 6$ inches fall in every 1200 feet (= $t$) is required to deliver $Q = 360$ cub. ft. of water per minute with uniform motion. Here we have given, with uniform motion, $h$, $l$, and $Q$, with the requirement that the section shall be trapezoidal, with $\theta = 45^\circ$, and of minimum frictional resistance. The following equations are available: Eq. of continuity $\ldots Q = Fv_0 \quad \ldots \quad (1)$ Eq. (8) preceding, for condition of least resistance $\therefore R' = \frac{1}{2}\sqrt{\frac{\sin{\theta}}{2 - \cos{\theta}}}F$. (9) From eq. (8), for uniform motion, $k = f'\frac{x'}{B}\frac{x'}{2g}$ $\ldots$ $\ldots$ (3') There are three unknown quantities, $v$, $F$, and $R'$. Solve 788 MECHANICS OF ENGINEERING. (1') for $v$; solve (2) for $F$; substitute their values in (3); whence $$\frac{h \sin \theta}{2} = \frac{2l}{2g} \cdot \frac{Q^2}{\sqrt{F}} \cdot v = \left[ \frac{2f + 2 - \cos \theta}{2gh \sin \theta} \right] F. \quad (4')$$ Since $f$ cannot be exactly computed in advance, for want of knowing the value of $R$, we calculate it approximately (eq. (6), § 544b) for an assumed value of $R$; insert it in the above equation, and thus find a new value of $F$; and then, from (8), a corresponding value of $v$, from which a new value of $f$ can be computed. Thus after one or two trials a satisfactory adjustment of dimensions can be secured. 545. Variable Motion.—If a steady flow of water of a delivery $Q_0 = P_0$, constant, takes place in a straight open channel the slope of whose bed has not the proper value to maintain a "steady motion," then the velocity will "ensue" (the flow is still steady, but variable); i.e., although the mean velocity in any one transverse section remains fixed (with lapse of time), this velocity has different values for different sections; but as the eq. of continuity, $$Q = Fv = F_0v_0 = F_1v_1,$$ still holds (since the flow is steady), the different sections have different areas. If, Fig. 616, a stream of water flows down an inclined trough without friction, the relation between the velocities $v_0$ and $v_1$ at any two sections 0 and 1 will be the same as for a material point sliding down a guide without friction (see § 79, latter part), viz.: $$\frac{v_1^*}{2g} = \frac{v_0^*}{2g} + h. \quad (f)$$ A diagram showing a stream of water flowing down an inclined trough. Fig. 616 VARIABLE MOTION. OPEN CHANNEL. 769 a.: equation of heads (really a case of Bernoulli's Theorem, § 493). But, considering friction on the bed, we must subtract the mean friction-head $f \frac{l}{R} \frac{w_a^*}{2g}$ [see eqs. (3) and (3'), § 548] lost between 0 and 1; this friction-head may also be written thus: $f \frac{l w_a^*}{F_+} \frac{w_a^*}{2g}$; and therefore eq. (1) becomes $$\frac{w_a^*}{2g} = \frac{w_a^*}{2g} + h - \frac{f l w_a^*}{F_+} \frac{w_a^*}{2g}, \quad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)$$ which is the formula for variable motion; and in it $l$ is the length of the section considered, which should be taken short enough to consider the surface straight between the end-sections, and the latter should differ but slightly in area. The subscript $m$ may be taken as referring to the section midway between ends, so that $w_m = w(0 + \frac{l}{2})$. The actual perimeter $w_m = w(0 + w_o)$, and $F_m = F(F_+ + F_-)$. Hence eq. (9) becomes $$h = \frac{w_m^*}{2g} - \frac{w_m^*}{2g} + \frac{f l(w_m + w_o)}{F_+ + F_-} \frac{w_m^*}{2g}; \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. \quad. (8)$$ and again, by putting $v_e = Q + F_-$, $v_i = Q + F_+$, we may write $$h = [\frac{1}{F_+} - 1\frac{1}{F_-} + 1\frac{f l(w_m + w_o)}{F_+ + F_-} (\frac{1}{F_+} + 1\frac{1}{F_-})]Q'_{-}; (4)$$ whence $$Q = -\sqrt{\frac{2gh}{F_+}}[\frac{1}{F_+} - 1\frac{1}{F_-} + 1\frac{f l(w_m + w_o)}{F_+ + F_-}] [\frac{1}{F_+} + 1\frac{1}{F_-}] (5)$$ From eq. (4), having given the desired shapes, areas, etc., of the end-sections and the volume of water, $Q$, to be carried per unit of time, we may compute the necessary fall, $h$, of the surface, in length $l = T$, while from eq. (8), having observed in an actual water-course the values of the sectional areas $F_+$ and $F_-$, the wanted perimeter $w_m$ and $w_o$, the lengths $d_+$ of the per- 770 MECHANICS OF ENGINEERING. tion considered, we may calculate $Q$ and thus gauge the stream approximately, without making any velocity measurements. As to the value of $f$, we compute it from eq. (6), § 542b, using for $R$ a mean between the values of the hydraulic radii of the end-sections. 544. Bands in an Open Channel.—According to Humphreys and Abbot's researches on the Mississippi River the loss of head due to a bend may be put $$h_b = \frac{v^2}{g} \cdot \frac{\delta}{\pi} \cdot \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \\ [equation] \\ h = f(\frac{v^2}{g}) + 6\frac{v^2\delta}{536\pi} - [ft.\ and.\ sec.] \\ [equation] \\ while eq. (2) of § 545 for variable motion would then become \\ [equation] \\ (e and $\delta$ as above.) (For "radian" see p. 544.) 547. Equations for Variable Motion, introducing the Depth. —Fig. 618. The slope of the bed being sin $\alpha$ (or simply $\alpha$, $\pi$ mean.), while that of the surface is different, viz., $$\sin\beta = s = h + l,$$ we may write $$h = d_c + l\sin\alpha - d_t,$$ Fig. 617. Fig. 618. VARIABLE MOTION. OPEN CHANNEL. 771 in which $d_s$ and $d_e$ are the depths at the end-sections of the portion considered (steady flow with variable motion). With these substitutions in eq. (4), § 545, we have, solving for $l$, $$l = \frac{d_s - d_e - \left( \frac{1}{F_s} + \frac{1}{F_e} \right) Q^2}{\frac{f w_a}{g} \left( \frac{1}{F_s} + \frac{1}{F_e} \right) Q^2 - \sin \alpha} \dots (1)$$ From which, knowing the slope of the bed and the shape and size of the end-sections, also the discharge $Q$, we may compute the length or distance, $l$, between two sections whose depths differ by an assigned amount $(d_s - d_e)$. But we cannot compute the change of depth for an assigned length $l$ from (6). However, if the width $b$ of the stream is constant, and the same at all depths; i.e., if all sections are rectangles having a common width; eq. (6) may be made simplified by introducing some approximations, as follows: We may put $$\left( \frac{1}{F_s} + \frac{1}{F_e} \right) Q = \frac{F_s^* - F_e^*}{F_s^* F_e^* g} = \frac{(F_s^* - F_e^*) (P_s + P_e)}{F_s^* F_e^* g} v_s^*$$ $$= \frac{(d_s - d_e)(d_s + d_e)}{d_s d_e} v_s^*, \text{ which approx. } = \frac{2(d_s - d_e)}{d_s d_e} v_s^*$$ and, similarly, $$\frac{w_a}{g} \left( \frac{1}{F_s} + \frac{1}{F_e} \right) Q = \frac{w_a(F_s^* + F_e^*) v_s^*}{(F_s + F_e) F_s^* F_e^* g}$$ which approx. = $\frac{w_a v_s^*}{d_b g}$ Hence by substitution in eq. (6) we have $$l = \frac{(d_s - d_e) \left[ 1 - \frac{2}{d_s d_e} v_s^* \right]}{\frac{f w_a v_s^*}{d_b g} - \sin \alpha} \dots (7)$$ § 547a. Backwater.—Let us suppose that a steady flow has been proceeding with uniform motion (i.e., the surface parallel 772 MECHANICS OF ENGINEERING. to the bed) in an open channel of indefinite extent, and that a vertical wall in now set up across the stream. The water rises and flows over the edge of the wall, or weir, and after a time a steady state is reached. We shall suppose that the depth of water close to the wall on the up-stream side is greater than $d_s$, the original depth. We now have "variable motion" above the weir, and at any distance $x$ up-stream from the weir the new depth $y$ is greater than $d_s$. This increase of depth is called backwater, and, though decreasing up-stream, may be perceptible several miles above the weir. Let $s$ be the slope of the original uniform motion (and also of present bed), and $v$ the velocity of the original uniform motion, and let $k = \frac{v^2}{g}$. Then, if the section of the stream is a shallow rectangle of constant width, we have the following relation (Rankine): $$x = \frac{1}{s}\left[y - y + (d_s - 2k(\phi - \phi_s)\right], \dots \dots (1)$$ where $\phi$ is a function of $\frac{y}{d_s}$, as per following table:
For $\frac{y}{d_s}$ 1.10 1.30 1.50 1.60 1.70
$\phi = \omega$ .80 .80 .876 .804 .718 .699
For $\frac{y}{d_s}$ 1.90 2.00 2.20 2.40 2.60 2.80
$\phi = .166$ .147 .123 .107 .099 .076 .065
.056
$\phi_s$ is found from $\frac{y_s}{d_s}$ precisely as $\phi$ from $\frac{y}{d_s}$, by use of the table. With this table and eq. (1), therefore, we can find $x$, the distance ("amplitude of backwater") from the weir of the point where any assigned depth $y$ (or "height of backwater", $y - d_s$) will be found. For example, Prof. Bowser cites the case from D'Aubuisson's Hydraulics of the river Weser in Germany, where the erection of a weir increased the depth at the weir from 9.5 ft. to 10 ft., the flow having been originally "uniform" for 10 miles. Three miles above the dam the increase $(y - d_s)$ of depth was 1.25 ft., and even at four miles it was 0.75 ft. CHAPTER VIII DYNAMICS OF GASEOUS FLUIDS. 548. Steady Flow of a Gas—[N.B. The student should now review § 499 up to eq. (5.)] The differential equation from which Bernoulli's Theorem was derived for any liquid, without friction, was [eq. (5), § 492] $$\frac{1}{g} v dv + ds + \frac{1}{y} dp = 0, \ldots \ldots \ldots (A)$$ and is equally applicable to the steady flow of a gaseous fluid, but with this difference in subsequent work, that the heaviness, $y$ ($§ 7$), of the gas passing different sections of the pipe or stream-line is, or may be, different (though always the same at a given point or section, since the flow is steady). For the present we neglect friction and consider the flow from a large receiver, whose great body of the gas is practically at rest, through an orifice in a thin plate, or a short nozzle with a rounded entrance. In the steady flow of a gas, since $y$ is different at different points, the equation of continuity takes the form Flow of weight per time unit = $F_v y_v = F_v y_s = \text{etc.}; \text{(a)}$ i.e., the weight of gas passing any section, of area $F$, per unit of time, is the same as for any other section, or $F y_v = \text{constant}, y$ being the heaviness at the section, and $v$ the velocity. 549. Flow through an Orifice—Remarks—in Fig. 619 we have a large rigid receiver containing gas at some tension, $p_s$, higher than that, $p_m$, of the (still) outside air (or gas), and at some absolute temperature $T_s$, and of some heaviness $y_s$; that is, in a state $n$. The small orifice of area $F$ being opened, the gas begins to escape, and if the receiver is very large, or if the supply is continually kept up by a blowing-engine, e.g., after 773 774 MECHANICS OF ENGINEERING. A very short time the flow becomes steady. Let $nm$ represent any stream-line (§ 405) of the flow. According to the ideal subdivision of this stream-line into laminae of equal mass or weight (not equal volumes), we have in the lape of time $dt$ establishing eq. (A) for any one lamina, each lamina in the lape of time $dt$ moves into the position just vacated by the lamina next in front, and assume precisely the same velocity, pressure, and volume (and therefore heaviness) as that from one had at the beginning of the $dt$. In its progress toward the orifice it expands in volume, its mass remains constant, its velocity, immovable at $n_1$, is gradually accelerated on account of the pressure from behind always being greater than that in front, until at $m$, in the "throat" of the jet, the velocity has become $v_m$, the pressure (i.e., tension) has fallen to a value $p_m$, and the heaviness has changed to $y_m$. The temperature $T_m$ (absolute) is less than $T_*$, since the expansion has been rapid, and does not depend on the temperature of the outside air or gas into which efflux takes place; though, of course, after the efflux has once free from tension, it will regain its original temperature in time. We assume the pressure $p_a$ (in throat of jet) to be equal to that of the outside medium (as was done with flow of water), so long as that outside tension is greater than $.527 p_a$ but if it is less than $.527 p_a$ and is even zero (a vacuum), experiment seems to show that $p_a$ remains equal to $.527$ of the interior tension $p_a$: probably on account of the expansion of the effluent gas beyond the throat, Fig. 620, so that although the tension in the outer edge, at $a$, of this jet is equal to that in outside medium, the tension at $m$ is greater because of the centripetal and centrifugal forces developed in the curved filaments between $a$ and $m$. (See § 553.) 550. Flow through an Orifice; Heaviness assumed Constant during Flow. The Water Formula.—If the inner tension $p_a$ ex- STeady Flow of Gases. 775 ceeds the outer, $p_m$, but slightly, we may assume that, like water, the gas remains of the same heaviness during flow. Then, for the simultaneous advance made by all the laminae of a stream-like Fig. 613, in the time $dt$, we may conceive an equation like eq. (A), for each lamina between $n$ and $m$, and corresponding terms added; i.e., (For reference) $\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots$d In general, $\gamma$ is different in the different laminae, but in the present case it is assumed to be the same in all; hence, with $m$ as datum level and $h$ = vertical distance from $n$ to $m$, we have, from eq. (B), $$\frac{v_n}{dy} - \frac{v_m}{dy} + 0 - h + \frac{p_n}{y} - \frac{p_m}{y} = 0.$$ But we may put $v_n = 0$; while $h$, even if several feet, is small compared with $\frac{p_n}{y} - \frac{p_m}{y}$ E.g., with $p_n = 15$ lbs. per sq. in. and $p_m = 16$ lbs. per sq. in., we have for atmospheric air at freezing temperature $$\frac{p_n}{y} - \frac{p_m}{y} = 1683 \text{ feet}.$$ Hence, putting $v_n = 0$ and $h = 0$ in eq. (1), we have $$\frac{v_m^*}{dy} = \frac{p_n - p_m}{y}.$$ (Water formula; for small difference of pressures, only.) The interior absolute temperature $T_n$ being known, the $v_n$ (interior heaviness) may be obtained from $v_n = p_n y_n T_n + T_{n,p}$ ($§ 472$), and the volume of flow per unit of time then obtained (first solving (2) for $v_n$) $$Q_a = F_a v_n,$$ where $F_a$ is the sectional area of the jet at $m$. If the mouth-piece or orifice has well-rounded interior edges, as in Fig. 541, A diagram showing a stream-like figure with various points labeled. 776 MECHANICS OF ENGINEERING. its sectional area $F$ may be taken as the area $F_m$. But if it is an orifice in "this plate," putting the coefficient of contraction = $C = 0.60$, we have $$F_m = OF = 0.60 F \quad \text{and} \quad Q_m = 0.60 Fv_m.$$ (4) This volume, $Q_m$, is that occupied by the flow per time-unit when in state $m$, and we have assumed that $y_m = y_s$; hence the weight of flow per time-unit is $$Q = Q_m y_m = F_m v_my_s = Fmv_y.$$ (5) **EXAMPLE.**—In the testing of a blowing-engine it is found capable of maintaining a pressure of 18 lbs. per sq. inch in a large receiver, from whose side a blast is steadily escaping through a "thin plate" orifice (circular) having an area $F = 4$ sq. inches. The interior temperature is 90° Cent. and the outlet tension 15 lbs. per sq. in. Required the discharge of air per second, both volume and weight. The data are: $p_a = 18$ lbs. per sq. in., $T_a = 303^\circ$ Abo. Cent., $F = 4$ sq. inches, and $p_w = 15$ lbs. per sq. in. Use ft.-lb.-sec. system. First, the heaviness in the receiver is $$y_s = \frac{p_a}{p_w} \cdot \frac{T_a}{T_s} \cdot y_t = \frac{18}{15} \times \frac{303}{14.7} \times .089 = .089 \text{ lbs. per cub. ft.}$$ Then, from eq. (2), $$v_m = \sqrt{\frac{2g p_a - p_w}{y_s}} = \sqrt{\frac{9 \times 32.2(144 \times 18 - 144 \times 15)}{0.089}} = 555.3$$ (97 per cent of this would be more correct on account of friction.) :. $Q_m = F_m v_my_s = Fmv_y \times 555.3 = 9.34$ cub. ft. per sec. at a tension of 15 lbs. per sq. in., and of heaviness (by hypothesis) = .089 lbs. per cub. ft. Hence weight $$G = 9.34 \times .089 = .82 \text{ lbs. per sec.}$$ FLOW OF GASES BY MARIOTTE'S LAW. 777 The theoretical power of the air-compressor or blowing-engine to maintain this steady flow can be computed as in Example 3, § 453. 551. Flow through an orifice on the Basis of Mariotte's Law ; or Isothermal Efflux. Since in reality the gas expands during flow through an orifice, and hence changes its heaviness (Fig. 619), we approximate more nearly to the truth in assuming this change of density to follow Mariotte's law, i.e., that the heaviness varies directly as the pressure, and thus imply that the force required to maintain the flow is proportional to $p$, and again integrate the terms of eq. $(B)$, but take care to note that, now, $y$ is variable (i.e., different in different lamina at the same instant), and hence express it in terms of the variable $p$ (from eq. $(2)$, § 475), thus: $$y = (y_+ + p_+ p)$$ Therefore the term $\int_{a}^{b} \frac{dp}{y}$ of eq. $(B)$ becomes $$\frac{p_+}{y_+} \int_{a}^{b} \frac{dp}{p} = \frac{p_+}{y_+} \log \frac{p_+}{p_+} \cdot \cdot \cdot (1)$$ and, integrating all the terms of eq. $(B)$, neglecting $h$, and calling $v_a$ zero, we have $$\frac{v_a^*}{2g} = \frac{p_+}{y_+} \log \frac{p_+}{p_+} - \text{"efflux by Mariotte's Law through orifice"} \cdot \cdot \cdot (3)$$ As before, $y_+ = \frac{p_+}{T_+ p_+}$, and the flow of volume per time-unit at $m$ is $$Q_m = F_m v_m; \cdot \cdot \cdot \cdot (3)$$ while if the orifice is in thin plate, $F_m$ may be put = .60 $F$, and the weight of the flow per time-unit = $G = F_m w_m$. (4) If the mouth-piece is rounded, $F_m = F =$ area of exit orifice of mouth-piece. A diagram showing a cross-section of an orifice with a stream of gas flowing through it. 778 MECHANICS OF ENGINEERING. EXAMPLE.—Applying eq. (3) to the data of the example in § 350, where $y_a$ was found to be 0.089 lbs. per cub. ft., we have $$e_m = \sqrt{2g\frac{p_a}{y_a}\log{\frac{p_a}{p_0}}}$$ $$= \sqrt{2 \times 32.2 \times \frac{18 \times 144}{0.089} \times 2.9025 \times \log{\frac{18}{15}}} = 584.7 \text{ ft. p. sec.}$$ :. $Q_a = F_av_a = 0.60 \times y_A \times 584.7 = 0.745$ cub. ft. per sec. Since the heaviness at $m$ is, from Mariotte's law, $$y_a = \frac{p_a}{p_0} y_a = \frac{18}{15} \times 0.089,$$ i.e., $y_a = -0.741$ lbs. per cub. ft., hence the weight of the discharge is $$G = Q_av_a = 0.745 \times (-0.741) = 0.729 \text{ lbs. per sec},$$ or about 12 per cent less than that given by the “ water formula.” If the difference between the inner and outer tensions had been less, the discrepancy between the results of the two methods would not have been so marked. § 552. Adiabatic Reflux from an Orifice.—It is most logical to assume that the expansion of the gas approaching the orifice, being rapid, is adiabatic (§ 478). Hence (especially when the difference between the inner and outer tensions is considerable) it is more accurate to assume $y$ as varying according to Poisson's Law, eq. (1), § 478; i.e., $y = [y_a + p_a]^{1/p}$, in integrating eq. (B). Then the term $$\int_{a}^{b} \frac{dp}{y} \text{ will } = \frac{p^A}{p_a} \int_{a}^{b} p^{-1} dp = \frac{p^A}{p_a} [p_b^A - p_a^A]$$ $$= -\frac{p_a}{y_a} [1 - (\frac{p_b}{p_a})^A];$$ ADIABATIC FLOW OF GASES THROUGH ORIFICES. 779 and eq. (B), neglecting $A$ as before, and with $v_a = 0$, becomes (See Fig. 619) $$\frac{v_a}{g} = \frac{3p_a}{y_a}\left[1 - \left(\frac{p_a}{p_0}\right)^{\frac{1}{k}}\right] \text{(Adiabatic flow; orifice).} \quad (1)$$ Having observed $p_a$ and $T_a$ in the reservoir, we compute $r_a = \frac{p_a y_a^{\frac{1}{k}}}{p_0}$ (from § 478). The gas at $m$, just leaving the orifice, having expanded adiabatically from the state $a$ to the state $m$, has cooled to a temperature $T_m$ (absolute) found thus (§ 475), $$T_m = T_a\left(\frac{p_a}{p_0}\right)^{\frac{1}{k}}, \quad \ldots \ldots \quad (2)$$ and is of a heaviness $$v_m = y_m\left(\frac{p_m}{p_0}\right)^{\frac{1}{k}}, \quad \ldots \ldots \quad (3)$$ and the flow per second occupies a volume (immediately on exit) $$Q_m = F_m v_m, \quad \ldots \ldots \quad (4)$$ and weighs $$G = F_m v_m g, \quad \ldots \ldots \quad (5)$$ EXAMPLE 1.—Let the interior conditions in the large reservoir of Fig. 619 be as follows (state $n$): $p_n = 224$ lbs. per sq. in., and $T_n = 294^\circ$ Abs. Cent. (i.e., 21st Cent.); while externally the tension is 15 lbs. per sq. inch, which may be taken as being $p_w =$ tension at $m$, the throat of jet. The opening is a circular orifice in "thin plate" and of one inch diameter. Required the weight of the discharge per second [ft., lb., sec.; $g = 32.2$]. First, $y_n = \frac{22.5 \times 144}{15} \times .0807 = .114$ lbs. per cub. ft. Then, from (1), $$v_n = \sqrt{\frac{3p_n}{y_n}\left[1 - \left(\frac{p_n}{p_0}\right)^{\frac{1}{k}}\right]}$$ $$= \sqrt{\frac{9 \times 32.2 \times 3 \times 22.5 \times 144}{0.114}\left[1 - .\overline{{v}}_{\frac{k}{k}}\right]} = 844 \text{ ft. per sec.}$$ 780 MECHANICS OF ENGINEERING. Now $$F = \frac{4\pi(1.5)^2}{6} = 0.0546 \text{ sq. ft.}$$ :. $$Q_m = CFv_m = .80 Fv_m = 0.60 \times 0.0546 \times 844 = 2.765$$ cub. ft. per sec., at a temperature of $$T_m = 294^\circ F = 257^\circ \text{ Abs. Cent.} = -16^\circ \text{ Cent.},$$ and of a heaviness $$y_m = 0.114 V(\frac{F}{H})^2 = 0.085 \text{ lbs. per cub. ft.},$$ so that the weight of flow per sec. $$G = Q_my_m = 2.765 \times .085 = .235 \text{ lbs. per sec.}$$ EXAMPLE 9.--Let us treat the example already solved by the two preceding approximate methods (§§ 550 and 551) by the present more accurate equation of adiabatic flow, eq. (1). The data were (Fig. 619): $$p_a = 18 \text{ lbs. per sq. in.; } T_a = 803^\circ \text{ Abs. Cent.; } p_h = 15 \text{ " " " " ; and } F = 4 \text{ sq. inches}$$ (F being the area of orifice). $$y_m$$ was found = .089 lbs. per cub. ft. in § 550; hence, from eq. (1), $$v_m = \sqrt{\frac{2 \times 32.2 \times 3 \times 18 \times 844}{0.089}[1 - V^2]} = 576.9 \text{ ft./per sec.}$$ From (4), $$Q_m = Fv_m = .6 Fv_m = .6 \times .089 \times 576.9 = 9.603 \text{ cub. ft. per sec;}$$ and since at m it is of a heaviness $$y_m = .089 V(\frac{F}{H})^2 = .0788 \text{ lbs. per cub. ft.,}$$ we have weight of flow per sec. $$= G = Q_my_m = 9.603 \times .0788 = 0.756 \text{ lbs. per sec.}$$ THEORETICAL MAXIMUM FLOW OF WEIGHT OF GAS. 781 Comparing the three methods for this problem, we see that By the "water formula," . . . $G = 0.82$ lbs. per sec. " isothermal formula, . . . $G = 0.729$ " " adiabatic formula, . . . $G = 0.756$" 553. Practical Notes. Theoretical Maximum Flow of Weight. —if in the equations of § 559 we write for brevity $p_w + p_a = w$ we derive, by substitution from (1) and (3) in (5), Weight of flow $= G Q_w Y_w = F_w \sqrt{g g_p Y_w} (1 - a_f) t$, . . . (1) per unit of time) This function of $x$ is of such a form as to be a maximum for $x = (p_w + p_a) = (\frac{1}{2}) = 512$; . . . (2) i.e., theoretically, if the state $n$ inside the reservoir remains the same, while the outside tension (considered $= p_w$ of jet, Fig. 619) is made to assume lower and lower values (so that $x = p_w + p_a$, diminishes in the same ratio), the maximum flow of weight per unit of time will occur when $p_w = 512$ $p_a$, a little more than half the inside tension. (With the more accurate value 1.41 (1.408), instead of 1.4, see § 478, we should obtain 327 instead of 512 for dry air; see § 549.) Prof. Cortotell says ("544 of his Applied Mechanics"): "The distinction of the theoretical divergence on limitation of the external pressure to which the limit given is an anomaly which had always been considered as requiring explanation, and M. St. Venant had already suggested that it could not actually occur. In 1866 Mr. R. D. Napier showed by experiment that the weight of steam of given pressure discharged from an orifice really is independent of the pressure of the medium into which efflux takes place *"; and in 1872 Mr. Wilson confirmed this result by experiments on the reaction of steam issuing from an orifice. *The explanation lies in the fact that the pressure in the When the difference between internal and external pressures is great, should be added. A diagram showing a water flow through a pipe with a valve. 789 MECHANICS OF ENGINEERING. centre of the contracted jet is not the same as that of the surrounding medium. The jet after passing the contracted section suddenly expands, and the change of direction of the fluid particles gives rise to centrifugal forces which cause the pressures to be greater in the centre of the contracted section than at the circumference; see Fig. 690. Prof. Cotterill then advises the assumption that $p_{a} = \frac{527}{p_{a}}$ (for air and perfect gases) as the mean tension in the jet at $m$ (Fig. 619), whereas the outside medium is at a tension less than $527 p_{a}$. He also says, "Contraction and friction must be allowed for, but these effects are so small in magnitude that the value of which, however, is more variable than that of the corresponding coefficient for an incompressible fluid. Little is certainly known on this point." See §§ 549 and 554. For air the velocity of this maximum flow of weight is $$\text{Val. of max. } G = \left[ \frac{997}{T_a T_t} \right] \text{ ft. per sec., . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . \sqrt{\frac{T_a}{T_t}} \right]$$ (3) where $T_a$ = abs. temp. in reservoir, and $T_t$ = that of freezing point. Rankine's Applied Mechanics (p. 584) mentions experiments of Drs. Joule and Thomson, in which the circular orifices used in the test of copper tubes of diameter 0.029 in., 0.035 in., and 0.084 in., while the outside temperature was about one half of that inside. The results were 84 per cent of those demanded by theory, a discrepancy due mainly, as Rankine says, to the fact that the actual area of the orifice was used in computation instead of the contracted section; i.e., contraction was neglected. 554. Coefficients of Efflux by Experiment For Orifices and Short Pipes. Small Difference of Tension.—Since the discharge through a short pipe from a reservoir is affected not only by contraction, but by slight friction at the edges, even with a rounded entrance, the theoretical results for the volume and weight of flow per unit of time in preceding paragraphs should be multiplied both by a coefficient of velocity $\phi$ and one for contraction $C$, as in the case of water; i.e., by a coefficient of efflux $\mu = \phi C$. (Of course, when there is no COEFFICIENTS OF EFFLUX. GAS. 783 contraction, $C = 1.00$, and then $\mu = \phi$ as with a well-rounded mouth-piece, (for instance, Fig. 541, and with short pipes.) Hence for practical results, with orifice and short pipes, we should write for the weight of flow per unit of time $$= G = \mu F_{\text{in}} y_s = \mu F \left(\frac{F_{\text{in}}}{F_{\text{out}}} \right) \sqrt{\frac{2g p_s p_c}{\rho g}} \left[1 - \left(\frac{y_s}{p_s}\right)^2 \right],$$ (1) (from the equations of § 555 for adiabatic flow, as most accurate; $p_s + p_c$ may range from $4$ to $1.00$. $F =$ area of orifices, or of discharging end of mouth-piece or short pipe; $y_s =$ heaviness of air in reservoir and $= T_s p_s + T_c p_c$, eq. (18) of § 437; and $\mu =$ the experimental coefficient of efflux. From his own experiments and those of Koch, D'Aubrisson, and others, Weibach recommends the following mean values of $\mu$ for various mouthpieces, when $p_s$ is not more than $4$ larger than $p_c$ (i.e., about $17 \%$ larger), for use in eq. (1): 1. For an orifice in a thin plate, . . . . . . . . . . $\mu = 0.56$ 2. For a short cylindrical pipe (inner corners not rounded), $\mu = 0.75$ 3. For a well-rounded mouth-piece (like that in Fig. 541), $\mu = 0.98$ 4. For a short conical convergent pipe (angle about $6^o$), $\mu = 0.92$ EXAMPLE.—(Data from Weibach's Mechanics.) "If the sum of the areas of two conical turrets of a blowing-machine is $F = 3$ sq. inches, the temperature in the reservoir $18^o$ Cent., the height of the attached (open) mercury manometer (see Fig. 464) 29 inches, and the height of the barometer in the external air 29 inches," we have (ft., lb., sec.) $$\frac{P_2}{P_1} = \frac{29}{29+8} = \frac{29}{37}, \quad T_s = 288^\circ \text{ Abs. Cent.;}$$ $$p_s = (\#) 14.7 \times 144 \text{ lbs. per sq. ft.;}$$ $$y_s = (\#) \times 0.0807 = 0.0816 \text{ lbs. per cub. ft.,}$$ while $$F' = \frac{T_s}{T_c} \text{ sq. ft. and (see above)} \mu = 0.92; \text{ hence}$$ $$\phi = 0.92 \times (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#) (\#).$$ 784 MECHANICS OF ENGINEERING. i.e., $G = 0.6076$ lbs. per second; which will occupy a volume $$V_1 = G + p_1 V = G + 0.0807 \times 7.59 = 7.93 \text{ cub. ft.}$$ at one atmosphere tension and freezing-point temperature, while at a temperature of $T_2 = 98^\circ$ Abs. Cent. and tension of $p_w = \frac{H}{H} \times 1$ of one atmosphere (i.e., in the state in which it was on entering the blowing-engine) it occupied a volume $$V = \frac{H}{H} \times 7.59 = 8.94 \text{ cub. ft.}$$ (This last is Weisbach's result, obtained by an approximate formula.) 555. Coefficients of Eflux for Orifices and Short Pipes for a Large Difference of Tension—For values $\frac{d}{D} > \frac{1}{2}$ and $< 2$, of the ratio $p_a : p_m$, of internal to external tension, Weisbach's experiments with circular orifices in thin plate, of diameters ($= d$) from 0.4 inches to 0.8 inches, gave the following results:
$\frac{d}{D}$ $p_a : p_m$ 1.00 1.00 1.40 1.65 1.90 2.00
diam. = 0.4, $\mu$ = .95 $p_a : p_m$ .55 .59 .66 .73 .76 .78
diam. = 0.6, $\mu$ = .90 $p_a : p_m$ .60 .67 .64 .68 .72 .78
Whence it appears that $\mu$ increases somewhat with the ratio of $p_a$ to $p_m$, and decreases slightly for increasing size of orifice. With short cylindrical pipes, internal edges not rounded, and three times as long as wide, Weisbach obtained $\mu$ as follows:































































































<
When the inner edges of the 0.4 in. pipe were slightly rounded, $\mu$ was found = .93; while a well-rounded mouth-piece of the form shown in Fig. 541 gave a value $\mu =$ from .965 to .985, for $p_a : p_m$ ranging from 1.35 to 2.00. These values of $\mu$ are use for use in eqs (1), above. 556. To find the Discharge when the Internal Pressure is measured in a Small Reservoir or Pipe, not much larger than the **VELOCITY OF APPROACH. GAES.** 753 Orifice.—Fig. 621. If the internal pressure $p_*$, and temperature $T_*$, must be measured in a small reservoir or pipe, $n$, whose sectional area $F_*$ is not very large compared with that of the orifice, $F$, (or of the jet, $F_{\infty}$,) the velocity $v_*$ at $a$ (velocity of approach) cannot be put to zero. Hence, in applying eq. $(B)$, § 550, to the successive laminar between $a$ and $m$, and integrating, we shall have, for adiabatic steady flow, $$\frac{v_*^2}{2g} - \frac{u_*^2}{2g} = \frac{3p_*}{y_*}\left[1-\left(\frac{p_*}{p_0}\right)^{\frac{1}{k}}\right] \dots \dots (1)$$ Instead of eq. (1) of § 552. But from the equation of continuity for steady flow of gases [eq. (a) of § 548], $F_*v_*y_* = F_{\infty}v_{\infty}y_{\infty};$ hence $v_* = \frac{F_*y_*}{F_{\infty}y_{\infty}}v_{\infty},$ while for an adiabatic change from $a$ to $m$, $\frac{y_*}{y_m} = \left(\frac{p_*}{p_m}\right);$ whence by substitution in (1), solving for $v_*$, we have $$v_* = \sqrt{\frac{3g\cdot\frac{3p_*}{y_*}\left[1-\left(\frac{p_*}{p_0}\right)^{\frac{1}{k}}\right]}{\sqrt{1-\left(\frac{F_*}{F_{\infty}}\right)\left(\frac{p_*}{p_0}\right)^{\frac{1}{k}}}}} \dots \dots (2)$$ As before, from §§ 472 and 478, $$y_* = \frac{p_*T_*}{P_*T_*}y_* \dots \dots (3)$$ and $$y_m = \left(\frac{p_m}{p_0}\right)y_m \dots \dots (4)$$ Having observed $p_*$, $p_m$, and $T_*$, then, and knowing the area $F$ of the orifice, we may compute $y_*$, $y_m$, and $v_*$, and finally the Weight of flow per time-unit = $G = \mu F_{\infty}v_{\infty}y_m,\dots\dots(5)$ 786 MECHANICS OF ENGINEERING. taking $\mu$ from § 554 or 555. In eq. (9) it must be remembered that for an orifice in "thin plate," $F_m$ is the sectional area of the contracted vein, and = $CF$, where $C$ may be put = $\frac{\mu}{\sqrt{2}}$. **Example.**—If the diameter of $AB$, Fig. 621, is 34 inches, and that of the orifice, well rounded, = 2 in., if $p_a = 1 \frac{1}{2}$ atmospheres = 1$\frac{1}{2} \times 14.7 \times 144$ lbs. per sq. ft., while $p_m = \frac{1}{2}$ of an atom, so that $\frac{P_m}{P_a} = \frac{1}{2}$, and $T_a = 283^\circ$ Abs. Cent., required the discharge per second, using the ft., lbs., and sec. From eq. (3), $$y_a = \frac{1}{2} \cdot \frac{H_0}{H_0} \times 0.0807 = 0.0433 \text{ lbs. per cub. ft.;}$$ whence (eq. (4)) $$y_m = (\frac{1}{2}) y_a = 0.07544 \text{ lbs. per cub. ft.}$$ Then, from eq. (3), $$e_m = \left[ \sqrt{64 \times 3^2 \times 15 \times 985 \times 144} - (\frac{1}{2}) H_0 \right] + \left[ \sqrt{(14)^2 \times (14)} - (\frac{1}{2}) H_0 \right]$$ $$= 558.1 \text{ ft. per sec.;}$$ $$\therefore G = 0.98 \left( \frac{1}{\sqrt{6}} \right) 558.1 \times 0.07544 = 9003 \text{ lbs. per sec.}$$ 557. Transmission of Compressed Air; through very Long Level Pipes. Steady Flow. CASE I. When the difference between the tensions in the reservoirs at the ends of the pipe is small.—Fig. 622. Under A diagram showing two reservoirs connected by a pipe with a valve in the middle. these circumstances it is simpler to employ the form of formula that would be obtained for a liquid by applying Bernoulli's Theorem, taking into account the "loss of head" occasioned TRANSMISSION OF COMPRESSED AIR. 787 by the friction on the sides of the pipe. Since the pipe is very long, and the change of pressure small, the mean velocity in the pipe, $v'$, seems to be nearly the same at all points along the pipe, and will not differ greatly between the velocity-heads at $n$ and $m$ will be neglected; a certain mean head $y'$ will be assigned to all the gas in the pipe, as if a liquid. Applying Bernoulli's Theorem, with friction, § 516, to the ends of the pipe, $n$ and $m$, we have (as for a liquid) $$\frac{p_n}{g} + \frac{p_m}{y'} + 0 = \frac{p'_n}{g} + \frac{p'_m}{y'} + 0 - 4f \frac{l}{d} \frac{v'^2}{g}. \quad (1)$$ Putting (as above mentioned) $v'_n - v'_m = 0$, we have, more simply, $$\frac{p_n - p_m}{y'} = 4f \frac{l}{d} \frac{v'^2}{g}. \quad (9)$$ The value of $f$ as coefficient of friction for air in long pipes is found to be somewhat smaller than for water; see next paragraph. 558. Transmission of Compressed Air. Experiments in the St. Gotthard Tunnel, 1878. — [See p. 96 of Vol. 24 (Feb. '81), Van Nostrand's Engineering Magazine.] In these experiments, the temperature and pressure of the flowing gas (air) were observed at each end of a long portion of the pipe which delivered the compressed air to the boring-machines three miles distant from the tunnel's mouth. The portion considered was selected at a distance from the entrance of the tunnel, to eliminate the fluctuating influence of the weather on the temperature of the flowing air. A steady flow being secured by proper regulation of the compressors and distributing tubes, observations were made of the pressure $(p)$ at internal compressor points as well as the external, at each end of a portion of pipe considered, and also at intermediate points; also of the weight of flow per second $Q = Q_y$, measured at the compressors under standard conditions ($0^\circ$ Cent. and one atmos. tension). Then knowing the $p$ and $T$ at any section of the pipe, the 788 MECHANICS OF ENGINEERING. heaviness $y$ of the air passing that section can be computed from $\frac{y}{y'} = \frac{P - P_0}{P_0}$ and the velocity $v$ of $G + F'y'$ being $\frac{L}{F'}$, being the sectional area at that point. Hence the mean velocity $v'$, and the mean heaviness $y'$, can be computed for this portion of the pipe whose diameter $d$ and length $l$. In the experiments cited it was found that at points not too near the tunnel-mouth the temperature inside the pipe was always about 3° Cent. lower than that of the tunnel. The values of $f$ in the different experiments were then computed from eq. (2) of the last paragraph; i.e., $$\frac{P_s - P_0}{y'} = 4f \cdot \frac{w''}{d} \cdot \frac{w''}{2g}$$ (3) all the other quantities having been either directly observed, or computed from observed quantities. THE ST. GOTHARD EXPERIMENTS. [Concrete quantities reduced to English units.]
$\frac{d}{D}$ $p_a : p_m$ 1.05 1.10 1.35 1.40 1.70 1.74
diam. = 0.4, $\mu$ = .78 $p_a : p_m$ .72 .77 .88
diam. = 0.6, $\mu$ = .95 $p_a : p_m$ .85
diam. = 1.0, $\mu$ = .95 $p_a : p_m$ 81
No. d (feet) (ft.) (inches) Air pressure, Pa Pa - P_0 Wm sq. ft. per sec. mean f
1 15093 0.4055 5.60 5.34 19.82 21.1 .0035
2 15092 0.4058 5.60 5.34 19.82 21.1 .0038
3 15092 0.4058 5.60 5.34 19.82 21.1 .0041
4 17113 .3765 5.34 5.00
























































































<
In the article referred to (Van Nostrand's Mag.) $f$ is not computed. The writer contented himself with showing that Weisbach's values (based on experiments with small pipes and high velocities) are much too great for the pipes in use in the tunnel. With small tubes an inch or less in diameter Weisbach found, for a velocity of about 80 ft. per second, $f=.0068$; for still higher velocities $f$ was smaller, approximately, in accordance with the relation $$f = \frac{.0542}{v'/(in ft. per sec)}$$ TRANSMISSION OF COMPRESSED AIR. 780 On p. 370, vol. xxiv, Van Nostrand's Mag., Prof. Robinson of Ohio mentions other experiments with large long pipes. From the St. Gothard experiments a value of $f=.004$ may be inferred for approximate results with pipes from 3 to 8 in. in diameter. EXAMPLE.--It is required to transmit, in steady flow, a supply of $G=297^{\circ}$ lbs. of atmospheric air per second through a pipe 30000 ft. in length (nearly six miles) from a reservoir where the tension is 6.0 atmos. to another where it is 5.8 atmos., the mean temperature in the pipe being $80^{\circ}$ Fahr., $=24^{\circ}$ Cent. (i.e. = $297^{\circ}$ Abs. Cent.). Required the proper diameter of pipe; $d=$ ? The value $f=.00425$ will be used, and the ft.-lb.sec. system of units. The mean volume passing per second in the pipe is $$Q=G+f' \cdot \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \\ Q=G+f' =\frac{Q}{f'}=\frac{Q}{4\pi d^2}.\quad(4)$$ The mean velocity may thus be written: $v'=\frac{Q}{f'}=\frac{Q}{4\pi d^2}$. (4) The mean heaviness of the flowing air, computed for a mean tension of 5.9 atmospheres, is, by § 472, $$y'=\frac{5.9\times14.7\times973}{1\times14.7\times297}\times0.807=0.431\,\mathrm{lbs.\ per\ cub.\ ft.;}$$ and hence, see eq. (8). $$Q'=\frac{G}{y'}=\frac{6.456}{0.431}=14.74\,\mathrm{cub.\ ft;}$$ at tension of 5.9 atmos., and temperature $297^{\circ}$ Abs. Cent. Now, from eq. (2), $$\frac{P_a-P_m}{y'}=\frac{4f'}{2g}\cdot\frac{l}{d}\left[\frac{Q'}{4\pi d^2}\right];$$ whence $$d'=\frac{4f'}{(4\pi)^3}\cdot\frac{y'}{(P_a-P_m)}\cdot\frac{Q'}{2g};\quad\quad(5)$$ 790 MECHANICS OF ENGINEERING. and hence, numerically, $$d = 4 \times 0.0485 \times 0.431 \times 30000 \times (14.7)^2 / (7.854)^2 (14.7 \times 144(6.0 - 5.80)) / 38.2 = 1.29 \text{ feet}.$$ 559. (Case II of § 557) Long Pipe, with Considerable Difference of Pressure at Extremities of the Pipe. Flow Steady.--Fig. 693. If the difference between the end-tensions is comparatively great, we can no longer deal with the whole of the air in the pipe at once, as regards ascribing to it a mean velocity and mean tension, but must consider the separate laminae, such as $AB$ (a short length of the air-stream) to which we may apply eq. (3) of § 557; $A$ and $B$ corresponding to the $n$ and $m$ of Fig. 622. Since $p_1$, $p_2$, $F_1$, $F_2$, and $v$ of § 557 correspond to the $-dy$, $dx$, $y$, and $v$ of the present case (short section or lamina), we may write $$\frac{dp}{y} = 4f \frac{\partial^2}{\partial y^2} ds.$$ .......................... (1) But if $G =$ weight of flow per unit of time, we have at any section, $F_{yy} = G$ (equation of continuity); i.e., $v = G + F_y,$ whence by substitution in eq. (1) we have $$\frac{dp}{y} = \frac{4f}{2g} Gds;$$ i.e., $-yd\!p = \frac{4fG}{2gF_{yy}} ds.$$ .......................... (2) Eq. (3) contains three variables, $y$, $p$, and $s$ (= distance of lamina from $n$). As to the dependence of the behaviour of $y$ on the tension $p$ in different laminar experiments shows that in most cases a uniform temperature is found to exist all along the pipe, if properly buried, or shaded from the sun; the loss of heat by adiabatic expansion being in great part made up by the heat generated by the friction against the walls of the GAS IN LONG PIPES. LARGE FALL OF TENSION. 791 pipe. This is due to the small loss of tension per unit of length of pipe as compared with that occurring in a short discharge pipe or nozzle. Hence we may treat the flow as tæothermal, and write $p + y = p_w + y_w$ (§ 475, Mariotte's Law). Hence $y = \frac{y_w}{p_w}$, which substituted into eq. (2) enables us to write: $$- pdp = \left[ \frac{A_f G' p_x}{2g F^2} \right] da \dots \dots \dots (3)'$$ $$\therefore - \int_0^l pdp = \left[ \frac{A_f G' p_x}{2g F^2} \right] \int_0^l da \dots \dots (3)$$ Performing the integration, noting that at $n' p = p_w$, $s = 0$, and at $m' p = p_w'$ and $s = l$, we have $$\frac{1}{2} [p_w' - p_w'] = \frac{4f l}{2gd} \cdot \frac{A_f G' p_x}{F^2} \cdot y_w' \dots \dots (4)$$ It is here assumed that the tension at the entrance of the pipe is practically equal to that in the head reservoir, and that at the end ($m'$) to that of the receiving reservoir; which is not strictly true, especially when the corners are not rounded. It will be remembered also that in establishing eq. (2) of § 557 (the basis of this problem), the effect of friction on the gas was neglected; i.e., the change of velocity in passing along the pipe. Hence eq. (4) should not be applied to cases where the pipe is so short, or the difference of end-tensions so great, as to create a considerable difference between the velocities at the two ends of the pipe. (See Addendum on p. 791.) EXAMPLE.—A well or reservoir supplies natural gas at a tension of $p_w = 30$ lbs. per sq. inch. Its heaviness at 0° Cent. and one atmosphere tension is 348 lbs. per cu. foot. In piping this gas through a long horizontal line, it is desired, a single four-inch pipe is to be employed, and the tension in the receiving reservoir (by proper regulation of the gas distributed from it) is to be kept equal to 16 lbs. per sq. in. (which would sustain a column of water about 3 ft. in height in an open water manometer, Fig. 465). 792 MECHANICS OF ENGINEERING. The mean temperature in the pipe being 17° Cent., required the amount (weight) of gas delivered per second, supposing leakage to be prevented (formerly a difficult matter in practice). Solve (4) for $G$, and we have $$G = \frac{4\pi d^2}{4T} \sqrt{\frac{p_0^2}{4T} + \frac{y_{ac}(p_a - p_w)}{p_w}}.$$ (5) First, from § 472, with $T_w = T_a = 290^\circ$ Abs. Cent., we compute $$P_w = P_a \cdot \frac{T_a}{T_w} = \frac{14.7 \times 144}{0.048} \cdot \frac{390}{278} = 6454 \text{ feet}.$$ Hence with $f = .005$, $$G = \frac{4\pi(d^2)^2}{4T} \sqrt{\frac{32.9 \times 144(30 \times 144)^2 - (16 \times 144)^2}{4 \times 0.05 \times 10560 \times 4645}}$$ = 0.337 lbs. per sec. (For compressed atmospheric air, under like conditions, we would have $G = 0.430$ lbs. per second.) Of course the proper choice of the coefficient $f$ has an important influence on the result. From the above result ($G = 0.337$ lbs. per second) we can compute the volume occupied by this quantity of gas in the receiving reservoir, using the relation $Q_w = \frac{G}{y_w}$. The heaviness $y_w$ of the gas in the receiving reservoir is most easily found from the relation $\frac{P_w}{y_w} = \frac{F_w}{y_a}$, which holds good since the flow is isothermal. I.e., $\frac{P_w}{y_w} = 6454$ ft.; whence $y_w = 0.049$ lbs. per cubic foot, $p_w$ being $16 \times 144$ lbs. per sq. ft. Hence $$Q_w = \frac{G}{y_w} = \frac{0.337}{0.049} = 6.794 \text{ cub. ft. per sec.}$$ FLOW OF GAS IN PIPES. 733 It should be said that the pressure at the up-stream end of the pipe depends upon the rate of flow allowed to take place. With no flow permitted, the pressure in the tube of a gas-well has in some cases reached the high figure of 500 or 600 lbs. per eq. in. 560. Rate of Decrease of Pressure along a Long Pipe.—Con- sidering further the case of the last paragraph, that of a straight, long, level pipe of uniform diameter, delivering gas from a storage reservoir into a receiving reservoir, we note that if in eq. (4) we retain $p_w$ to indicate the tension at the receiving reservoir, but let $p_w'$ denote in turn the tension at points in the pipe successively further and further (a distance $x$ from the receiving reservoir), we may write eq. (5) for $x$ and obtain the equation (between two variables $p_w$ and $x$) $$p_w - p_w' = \text{Const.} \times x.$$ (6) This can be used to bring out an interesting relation men- tioned by a writer in the Engineering News of July 1887 (p. 71), viz., the fact that in the parts of the pipe more distant from the receiving end, $m$, the distance along the pipe in which a given loss of pressure occurs is much greater than near the receiving end. To make a numerical illustration, let us suppose that the pipe is of such size, in connection with other circumstances, that the tension $p_w$ at $A$, a distance $x$ six miles from $m$, is two atmospheres; that is, that the receiving reservoir being one atmosphere; that is, that the loss of tension between $A$ and $m$ is one atmosphere. If we express tensions in atmospheres and distances in miles, we have for the value of the constant in eq. (6), for this case, $$\text{Const.} = (4 - 1) + 6 = 9; \quad (\text{for assumed units}) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)$$ Now let $p_w' = \text{the tension at } B$, a point 18 miles from $m$, and we have, from eqs. (6) and (7), the tension at $B$ = 3.16 atmospheres. Proceeding in this manner, the following set of values is obtained:
$x$ $p_w - p_w'$
0 2
6 2
12 2
18 2
A table showing values of \( p_w - p_w' \) for different values of \( x \). The table shows: - For \( x = 0 \), \( p_w - p_w' = 2 \). - For \( x = 6 \), \( p_w - p_w' = 2 \). - For \( x = 12 \), \( p_w - p_w' = 2 \). - For \( x = 18 \), \( p_w - p_w' = 2 \). 794 MECHANICS OF ENGINEERING.
Point Total distance from m. Distance between two points Tension at point Loss of tension in mm.
P 128 miles 96 miles 8.00 atm. 1.25 atm.
R 96 30 6.75 1.30
D 64 30 6.75 1.30
C 32 18 4.35 1.19
A 16 18 4.35 1.19
B 8 6 2.00 1.00
m' 0"
If the distances and tensions in the second and fourth columns be plotted as abscisses and ordinates of a curve, the latter is a parabola with its axis following the axis of the pipe: its vertex is not at $m'$, however. 561. Long Pipe of Variable Diameter.—Another way of stating the fact mentioned in the last paragraph is as follows : At the up-stream end of the pipe of uniform diameter the gas is of much greater density than at the other extremity (the heaviness is directly as the tension, the temperature being assumed the same throughout the pipe), and the velocity of its motion is smaller than at the discharging end (in the same ratio). It is true that the frictional resistance per unit of length of pipe varies directly as the heaviness [eq. (1), § 610], but also true that it varies as the square of the velocity; so that, for instance, if the pressure at a point A is double that at B in the pipe of uniform diameter, it will be seen that the heaviness at A is double and half, respectively, those at B, and thus the gas at A is subjected to only half the frictional resisting force per foot of length as compared with that at B. Hence the relatively small diminution, per unit of length, in the tension at the up-stream end in the example of the last paragraph. In the pipe of uniform diameter, as we have seen, the greater part of the length is subjected to a comparatively high heavi- ness, and it is thus under a greater liability to lose by leakage than if it consisted of several sections of unequal diameter. The total "hoop-tension" ($\frac{4}{3}\pi$) in a unit length of pipe, also, is proportional to the gas tension, and thinner walls might be employed for the down-stream portions of the pipe if the gas FLOW OF GAS IN PIPES. 705 tension in those portions could be made smaller than as shown in the preceding example. To secure a more rapid fall of pressure at the up-stream end of the pipe, and at the same time provide for the same delivery of gas as with a pipe of uniform diameter throughout, a pipe of variable diameter may be employed, that diameter being considerably smaller at the inlet than that of the uniform pipe but progressively enlarging down-stream. This will require the diameters to be greater at the inlet than in the uniform pipe, and if the same thickness of metal were necessary throughout, there would be no saving of metal, but rather the reverse, as will be seen ; but the diminished thickness made practicable in those parts from a less total hoop tension than in the corresponding parts of the uniform pipe more than compensates for the extra metal due to increased circumference, aside from the diminished liability to leakage, which is of equal importance. A simple numerical example will illustrate the foregoing. The pipe being circular we may replace $F$ by $\frac{1}{4\pi d^2}$ in equation (4), and finally derive, $G$ being given, $$d = \text{Const} \times \left[\frac{1}{P_2 - P_1}\right]^{1/4} = C \cdot \left[\frac{1}{P_2' - P_1'}\right]^{1/4}. \quad (8)$$ Let $A$ be the head reservoir, and $m'$ the receiving reservoir, and $B$ a point half-way between. At $A$ the tension is 10 atmospheres; at $m'$, 2 atmospheres. For transmitting a given weight of gas per second, through a pipe of constant diameter throughout, that diameter must be (tensions in atmospheres) $\frac{9}{2}$ (being the length), by eq. (8), $$d = CI_A \left[\frac{9}{100 - \frac{9}{2}}\right]^{1/4} = CI_A(0.028)^{1/4} = 0.48 CI_A. \quad (8).$$ If we substitute for the pipe mentioned, another having a constant diameter $d$, from $A$ to $B$, where we wish the tension to be 5 atmospheres, and a different constant diameter $d'$, from $B$ to $m'$, we derive similarly, $$d' = CI_A \left[\frac{1}{100 - \frac{25}{2}}\right]^{1/4} = 0.42 CI_A.$$ 706 MECHANICS OF ENGINEERING. and $$d_s = C \left[ \frac{1}{25 - 4} \right]^{1/3} = 0.54 \, C D$$ It is now to be noted that the sum of $d_s$ and $d_t$ is slightly greater than the double of $d_s$; so that if the same thickness of metal were used in both designs the compound pipe would require a little more material than the uniform pipe; but, from the reasoning given at the beginning of this paragraph, that thing will not be much considerably less in the downstream part of the compound pipe than that which is saved. In case of a cessation of the flow, the gas tension in the whole pipe might rise to an equality with that of the head-reservoir were it not for the insertion, at intervals, of automatic regulators, each of which prevents the decrease of tension on its down-stream side below a fixed value. To provide for changes of length due to rise and fall of temperature, the pipe is laid with slight undulations. It is also a time-saving conclusion that a given pipe of variable diameter connecting two reservoirs of gas at specified pressures will deliver the same weight of gas as before, if turned end for end. This follows from equation (3)'s § 559. With $d$ variable, (3) becomes (with $F = 1 + d^2$) $$\int_{-pdp}^{pdp} G^* C^* \int_{d^2}^{pdp} \frac{ds}{d^2}; \quad i.e., \quad G^* = \frac{pd^2 - pdp}{2C^* \int_{d^2}^{pdp} \frac{ds}{d^2}}.$$ (9) ($G^*$ is a constant.) But $\int_{-pdp}^{pdp} \frac{ds}{d^2}$ is evidently the same in value if the pipe be turned end for end. In commenting on this circumstance, we should remember (see § 559) that the loss of pressure along the pipe is ascribed entirely to frictional resistance, and in no degree to changes of velocity (inertia). On p. 73 of the *Engineering* for July 1887 are given the full particulars of a design of a compound pipe in actual use, and delivering natural gas. The pressure in the head-reservoir is 310 lbs. per sq. in.; that in the receiving reservoir, 65. For 2.84 miles from the head-reservoir the diameter of the pipe is NATURAL GAS. COEFFICIENT OF FLUID FRICTION. 797 8 in.; throughout the next 2.75 miles, 10 in.; while in the remaining 3.84 miles the diameter is 18 in. At the two points of junction the pressures are stated to be 185 and 132 lbs. per sq. in., respectively, during the flow of gas under the conditions mentioned. 561a. Values of the Coefficient of Fluid Friction for Natural Gas.—In the Ohio Report on Economic Geology for 1888 may be found an article by Prof. S. W. Robinson of the University of that State describing a series of interesting experiments made by him on the flow of natural gas from orifices and through pipes. By the insertion of Pitot tubes approximate measurements were made of the velocity of the stream of gas in a pipe. The following are some of the results of these experiments, $p_1 - p_0$, representing the loss of pressure (in lbs. per sq. inch) per mile of pipe-length, and $f$ the coefficient of fluid friction, in experiments with a six-inch pipe:
$p_1 - p_0$ 1.00 1.50 2.00 2.50 3.00 3.50
$f$ 0.0025 0.0037 0.0050 0.0059 0.0070 0.0086
In the flow under observation Prof. Robinson concluded that $f$ could be taken as approximately proportional to the fourth root of the cube of the velocity of flow; though calling attention to the fact that very reliable results could hardly be expected under the circumstances. 561b. (Addendum.) Theothermal Flow in a Level Pipe, with Consideration of Inertia.—In eq. (1) of p. 697 neglect $d\sigma$, put $\omega + F = 4 + d$, and divide through by $\omega^2$. In the second term put $G^2 + F^2y^2$ for $\omega^2$ and then $p(y_n + p_n)$ for $y$. We now find the variables separated, and an integration for steady flow obtain (after putting $v_n + v_n = Fy_n + Fy_n = p_n + p_n$), $$\log(\frac{p_n}{p_m}) - \frac{g(Fx_n - px_n)}{2} \cdot \frac{F^2y_n}{G^2} = \frac{3pl}{D}$$ [Notation as in § 837 with $O = Fr.$.] A diagram showing a level pipe with a valve at one end and a pressure gauge at the other end. CHAPTER IX. IMPULSE AND RESISTANCE OF FLUIDS. 502. The so-called "Reaction" of a Jet of Water flowing from a Vessel.--In Fig. 634, if a frictionless but water-tight plug $B$ be inserted in an orifice in the vertical side of a vessel mounted on wheels, the resultant action of the water on the rigid vessel (as a whole) consists of its weight $G$, and a force $P' = Fhy$ (in which $F$ is the area of orifice) which is exerted by the hydrostatic pressures on the vessel wall toward the right ($\parallel$ to paper) over those toward the left, since the pressure $P' = Fhy$, exerted on the plug is felt by the post $C$, and not by the vessel. Hence the post $D$ receives a pressure $$P' = Fhy.$$ .......................... (1) Let the plug $B$ be removed. A steady flow is then set up through the orifice, and now the pressure against the post $D$ is $2Fhy$ (as will be shown in the next paragraph); it is not only this pressure $Fhy$ lacking at the orifice, but the sum of all the horizontal components ($\parallel$ to paper) of the pressures of the liquid filaments against the vessel wall around the orifice is less than its value before the flow began, by an amount $Fhy$. A resistance $R = 2Fhy$ being provided, and the post removed, a slow uniform motion may be maintained toward the right, the working force being $2Fhy = P'$. Fig. 634. **"REACTION" OF A JET.** 799 (see Fig. 635 ; $R$ is not shown). If an insufficient resistance be furnished before removing the post $D$, the vessel will begin to move toward the right with an acceleration, which will disturb the surface of the water and change the value of the horizontal force. This force $P' = 2F\gamma h \dots (2)$ is called the "reaction" of the water-jet; $\gamma$ is the heaviness of the liquid (q. p.). Of course, when the vessel moves, the water level sinks and the " reaction" diminishes accordingly. Looked upon as a motor, the vessel may be considered to be a piston-less and valve-less water-pressure engine, carrying its own reservoir with it. In Case II of § 500 we have already had a treatment of the "Reaction-wheel" or " Barker's mill," which is a practical machine operating on this principle, and will be again considered in "Notes on Hydraulic Motors." 583. " Reaction" of a Liquid Jet on the Vessel from which it issues.—Instead of showing that the pressures on the vessel close to the orifice are less than they were when there was no flow by an amount $F\gamma h$ (a rather lengthy demonstration), another result will be given, of greater simplicity but somewhat fanciful. If a man standing on the rear platform of a car is to take up in succession, from a basket on the car, a number of balls of equal mass $M$, and project each one in turn horizontally backward with an acceleration $p$, can accomplish this only by exerting against each ball a pressure $= M p$, and in the opposite direction against the car an equal pressure $= M p$. If this action is kept up continuously the car is subjected to a constant and continuous forward force $= F \gamma h$. Similarly, the backward propulsion of the jet of water in the case of the vessel at rest must occasion a forward force against the vessel of a value dependent on the fact that in each small interval of time $\Delta t$ a small mass $dM$ of liquid has its velocity changed from zero to a backward velocity of $v = V^2 \frac{dM}{h}$; that A diagram showing a water jet exiting from a pipe into a container. The container is labeled with arrows indicating movement to the right. 800 MECHANICS OF ENGINEERING. is, has been projected with a mean acceleration of $p = \frac{v - 0}{dt}$, so that the forward force against the vessel is $P'' = \text{mass} \times \text{acc.} = \frac{dM.v}{dt}$ . . . . (3) If $Q$ is the volume of water discharged per unit time, then $dM = \frac{Qy}{g} dt$, and since also $Q = Fv = F\sqrt{2gh}$, eq. (3) becomes "Reaction" of jet $= P'' = 2Fh.y$. . . . (4) (A similar proof, resulting in the same value for $P''$, is easily made if the vessel has a uniform motion with water surface horizontal.) If the orifice is in "thin plate," we understand by $F$ the area of the contracted section. Practically, we have $v = \phi\sqrt{2gh}$ (§ 495), and hence (3) reduces to $P'' = 2\phi^2Fhy$. . . . . . . . . . . . . (5) Weissbach mentions the experiments of Mr. Peter Ewart of Manchester, England, as giving the result $P'' = 1.73Fhy$ with a well-rounded orifice as in Fig. 625. He also found $\phi = .94$ for the same orifice, so that by eq. (4) we should have $P'' = 9(.94)^2Fhy = 1.71Fhy$. With an orifice in thin plate Mr. Ewart found $P'' = 1.14Fhy$. As for a result from eq. (4), we must put, for $F$, the area of the contracted section $64F$ (§ 495), which, with $\phi = .96$, gives $P'' = 9(.96)^2 \cdot 64Fhy = 1.18Fhy$. . . . (6) Evidently both results agree well with experiment. Experiments made by Prof. J. B. Webb at the Stevens Institute (see Journal of the Franklin Inst., Jan., '88, p. 35) also confirm the foregoing results. In these experiments the vessel was suspended on springe and the jet directly downwards, so that the "reaction" consisted of a diminution of the tension of the springe during the flow. 504 Impulse of a Jet of Water on a Fixed Curved Vane (with Borders).—The jet passes tangentially upon the vane. Fig. IMPULSE OF JET. 801 696. $B$ is the stationary nozzle from which a jet of water of cross-section $F$ (area) and velocity = $c$ impinges tangentially upon the vane, which has plane borders, parallel to paper, to prevent the lateral escape of the jet. The curve of the vane is not circular necessarily. The vane being smooth, the velocity of the water in its curved path remains $= c$ at all points along the curve. Conceive the curve divided into a great number of small lengths each $= dc$, and subtending some angle $\alpha = d\phi$ from its own centre of curvature, its radius of curvature being $r$ (different for different $dc$), which makes some angle $\phi$ with the axis $\Gamma$ ($\gamma$ to original straight jet $BA$). At any instant of time there is an arc of water $AD$ in contact with the vane, exerting pressure upon it. The pressure $dP$ of any $dc$ of the vane against the small mass of water $Fd\phi$, $\gamma + g$ then in contact with it is the "deviating" or "centripetal" force accountable for its motion in a curve of radius $r$, and hence must have a value $$dP = \frac{Fdyds}{g} \cdot \frac{e^{\prime}}{r} \cdot \ldots \ldots (76) \ldots (1)$$ The opposite and equal of this force is the $dP$ shown in Fig. 614, and is the impulse or pressure of this small mass against the vane. Its $X$-component is $dX = dP \cdot \sin \phi$. By making $\phi$ vary from 0 to $\alpha$, and adding up the corresponding values of $dX$, we obtain the sum of the $X$-components of the small pressures exerted simultaneously against the vane by the arc of water then in contact with it; i.e., noting that $de = rd\phi$ $$\therefore \int_{0}^{\alpha} dX = \int_{0}^{\alpha} dP \cdot \sin \phi = \frac{Fyds}{g} \int_{0}^{\alpha} \sin \phi \cdot r ds$$ $$= \frac{Fyds}{g} \int_{0}^{\alpha} [\sin \phi]_0^{\alpha} ds = \frac{Fyc'ds}{g} - \cos \varphi;$$ 809 MECHANICS OF ENGINEERING. hence the $X$-impulse $\{ = \frac{Fy'}{g} [1-\cos\alpha] = Qyc \frac{[1-\cos\alpha]}{g}, (3) \}$ in which $Q = F_0 = \text{volume of water which passes through the nozzle (and also = that passing over the vane, in this case)}$ per unit of time, and $\alpha = \text{angle between the direction of the stream leaving the vane (i.e., at } D\text{) and its original direction (BA of the jet); i.e., } \alpha = \text{total angle of deviation. Similarly, the sum of the } Y\text{-components of the } dP'\text{'s of Fig. 636 may be shown to be}\$ $Y\text{-impulse on fixed vane} = \int_0^{\infty} u'P' \cdot \cos\phi = \frac{Qyc}{g} \sin\alpha - (3')$ Hence the resultant impulse on the vane is a force $P'' = \sqrt{X'^2 + Y'^2} = \frac{Qyc}{g} \sqrt{2(1-\cos\alpha)}, . . . (3)$ and makes such an angle $\alpha'$, Fig. 627, with the direction BA, that $\tan\alpha' = \frac{Y'}{X'} = \frac{\sin\alpha}{1-\cos\alpha}. . . . . . . . . . . . . . . . . . . . . . . (4)$ For example, if $\alpha = 90^\circ$, then $\alpha' = 45^\circ$; while if $\alpha = 180^\circ$, Fig. 638, we have $\alpha' = 0^\circ$; i.e., $P''$ is parallel to the jet BA, and its value is $P'' = 9Q'y'\frac{c}{g}.$ IMPULSE OF JET ON VANE. 803 585. Impulse of a Jet on a Fixed Solid of Revolution whose Axis is Parallel to the Jet—If the curved vane, with borders, of the preceding paragraph be replaced by a solid of revolution, Fig. 629, with its axis in line with the jet, then the pressure of the jet upon it will simply be the sum of the $X$-components (i.e., = to $BA$) of the pressures on all elements of the surface at a given instant; i.e., $$X = P^{\prime} = QY \frac{c}{g}(1 - \cos\alpha); \quad \ldots \quad (5)$$ while the components $\gamma$ to $X$, all directed toward the axis of the solid, neutralize each other. For a flat plate, then, Fig. 630, at right angles to the jet, we have for the force, or "impulse" (with $\alpha = 90^\circ$), $$P^{\prime\prime} = QY \frac{c}{g} + FZ \gamma = 2F \frac{c}{g} \gamma. \quad (6)$$ The experiments of Bidone, made in 1838, confirm the truth of eq. (6) quite closely, as do also those of two students of the University of Pennsylvania at Philadelphia (see Jour. of the Frank. Inst. for Oct. '87, p. 258). Eq.(6) is applicable to the theory of Pitot's Tube (see § 539), Fig. 631, if we consider the edge of the tube plane as quite wide. The water in the tube at $A$ and its section at $A$ (of area $= F$) may be treated as a flat vertical plate receiving not only the hydrostatic pressure $FZ$, but to the depth $y$ below the surface, but a continuous impulse $P^{\prime\prime} = FZ\gamma + g$ [see eq. (6)] Fig. 629. Fig. 630. Fig. 631. 804 MECHANICS OF ENGINEERING. For the equilibrium of the end $A$, of the stationary column $AD$, we must have, therefore, $$F_{ay} + \frac{F_0y}{g} = F_{ay} + FA_y; \text{ i.e., } A' = (3.0) \frac{c'}{2g}. . . (7')$$ The relation in equation (7) corresponds reasonably well with the results of Weisbach's experiments with the instrument mentioned in § 539. Piot himself, on trial of an instrument in which the edges of the tube at $A$ were made flaring or conically divergent, like a funnel, found $$h' = (1.5) \frac{c'}{2g}; . . . . . . . . . . . . . . . (7)'$$ while Darcy, desires that the end of the tube should occasion as little disturbance as possible in the surrounding stream, made the extremity small and conically convergent.* The latter obtained the relation $$h' = \text{ almost exactly } (1.0) \frac{c'}{2g}. . . . . . . . . (7)'$$ (See § 539.) If the solid of revolution is made cup-shaped as in Fig. 632, we have (as in Fig. 629) $\alpha = 180^\circ$, and therefore, from eq. (5), $$P'' = 2Q_0\frac{c}{g} = \frac{2F_0y}{g} = 4F(\frac{c'}{2g})y; . . . . . . (8)$$ Fig. 632. EXAMPLE.—Fig. 632. If $c = 30$ ft. per sec. and the jet (cylindrical) has a diameter of 1 inch, the liquid being water, so that $y = 62.5$ lbs. per cub. ft., we have $(ft., lb., sec.)$ the impulse (force) $P'' = \frac{2 \times 1^{\prime}\frac{1}{12} \times 900 \times 62.5}{32.2} = 19.05$ lbs. Experiment would probably show a smaller result. * See p. 888 for Mr. Freeman's Experiments. IMPULSE OF JET ON MOVING VANE. 805 566. Impulse of a Liquid Jet upon a Moving Vane having Lateral Borders and Moving in the Direction of the Jet.—Fig. 633. The vane has a motion of translation (§ 108) in the same direction as the jet. Call this the axis $X$. It is moving with a velocity $v$ away from the jet (or, if toward the jet, $v$ is negative). We consider $v$ to be its acceleration being prevented by a proper resistance (such as a weight = $G$) to balance the $X$-components of the are-presures. Before coming in contact with the vane, which it does tangentially (to avoid rotation), the absolute velocity ($\S$ 83) of the water in the jet = $c$, while its velocity relatively to the vane at $A$ is $c - v$; and it will now be proved that the relative velocity along the vane is constant. See Fig. 634. Let $u$ be the velocity of the vane (of each point of it, since its motion is one of translation), and $w$ be the velocity of a water particle (or small mass of water of length $=da$) relatively to the point on the vane at which it is situated. Then, the absolute velocity of the small mass, is the diagonal formed on $u$ and $v$. Neglecting friction, the only actual force acting on the mass is $P$, the pressure of the vane against it, and this is normal to the curve. Now an imaginary system of forces, equivalent to this actual system of one force $P$, i.e., capable of producing the same motion in the mass, may be conceived of, consisting of the individual forces which would produce, separately, the separate motions of which the actual motion of this small mass $M$ is compounded. These components must satisfy: 1. A horizontal uniform motion of constant velocity = $v$; and 2. A motion in the arc of a circle of radius = $r$ and with a Fig. 633. 633 806 MECHANICS OF ENGINEERING. velocity = u, which we shall consider variable until proved otherwise. Motion 1 is of such a nature as to call for no force (by New- ton's first law of motion), while motion 2 could be maintained by a system of two forces, one normal, $P_n = \frac{Mdu}{r}$, and the other tangential, $P_t = M\frac{du}{dt}$ [see eq. (5), p. 76]. This imaginary system of forces is shown at (IL), Fig. 634, and is equiv- ![Diagram showing two forces acting on a body: one normal and one tangential.](image) alent to the actual system at (L). Therefore $\mathbb{Z}$(tang. compo.) in (L) should be equal to $\mathbb{Z}$(tang. compo.) in (IL); whence we have $$P_t = 0; \text{ i.e., } M\frac{du}{dt} = 0; \text{ or } \frac{du}{dt} = 0; . . .$$ (1) i.e., $u$ is constant along the vane and is equal to $c - v$ at every point. (The weight of the mass has been neglected since the height of the vane is small.) In Fig. 634 the symbol $w_o$ has been used instead of $c$, and the point 0 corresponds to A in Fig. 633. [N.B.] If the motion of the vane were rotary, about an axis to $A.P$ (or to $c$), this relative velocity would be different at different points. See Notes on Hydraulic Motors. If the radius of motion of the point $A$, however, is quite large compared with the projection of $AD$ upon this radius, the relative velocity is approximately $= c - v$ at all parts of the vane, and will be taken $= c - v$ in treating the "Hardy gurdy" in § 567.] A diagram showing two forces acting on a body: one normal and one tangential. WORK OF JET ON VANE. S07 By putting $X$ (normal component) of $(L) = X$ (normal com- ponent) in (II) we have $$P = P_x; \quad i.e., \quad P = M\frac{u^2}{r} = \frac{M(c-v)^2}{r}; \quad . . . \quad (3)$$ so that to find the sum of the $X$-components of the pressures exerted against the vane simultaneously by all the small masses of water in contact with it at any instant, the analysis differs from that in § 564 only in replacing the $c$ of that article by the $(c-v)$ of this. Therefore $$X(X_{pressure}) = P_x = \frac{F_y}{g}(c-v)^2[1-\cos a], \quad . . . \quad (3)$$ (where $a$ is the angle of total deviation, relatively to vane, of the stream leaving the vane, from its original direction), and is seen to be proportional to the square of the relative velocity. $F$ is the sectional area of jet, and $\gamma$ the heaviness ($\beta$) of the liquid. The $Y$-component (or $P_y$) of the resultant impulse is counteracted by the support $EF$, Fig. 638. Hence, for a uniform motion to be maintained, with a given velocity $v$, the weight $G$ must be made $= P_x$ of eq. (3). (We here neglect friction and suppose the jet to preserve a practically horizontal direction for an indefinite distance before meeting the vane. If this uniform motion is to be toward the jet, $v$ will be negative; if away from it, $v$ will be positive.) For a positive $v$ of same numerical value, As to the doing of work [§ 138, etc.], or exchange of energy, between the two bodies, jet and vane, during a uni- form motion away from the jet, $P_x$ exerts a power of $$L = P_xv = \frac{F_y}{g}(c-v)^2[1-\cos a], \quad . . . \quad (4)$$ in which $L$ denotes the number of units of work done per unit of time by $P_x$; i.e., the power ($\beta$ 130) exerted by $P_x$. If $v$ is negative, call it $-v$, and we have the Power expended } \\ {by vane upon jet } &= P_xv = \frac{F_y}{g}(c+v)^2[1-\cos a]. \quad . . . \quad (5)\\ 808 MECHANICS OF ENGINEERING. Of course, practically, we are more concerned with eq. (4) than with (5). The power $L$ in (4) is a maximum for $v = 4c$; but in practice, since a single moving vane or float cannot utilize the water of the jet as fast as it flows from the nozzle, let us conceive of a succession of vanes coming into position consecutively, so that each one has at first the same velocity $v$; then the portion of jet intercepted between two vanes is at liberty to finish its work on the front vane, while additional work is being done on the hinder one; i.e., the water will be utilized as fast as it issues from the nozzle. With such a series of vanes, then, we may put $Q' = P_c$, the volume of flow per unit of time from the nozzle, in place of $P(c - v)$ = the volume of flow per unit of time over the vane, in eq. (4); whence $$\text{Power exerted on } \frac{Q'}{g} = L' = \frac{Q'}{g}(1 - \cos \alpha)(c - v)v. \quad (6)$$ Making $v$ variable, and putting $dL' + dv = 0$, whence $c - 2v = 0$, we find that for $v = 4c$, $L'$, the power, is a maximum. Assuming different values for $\alpha$, we find that for $\alpha = 180^\circ$, i.e., by the use of a semicircular vane, or of a hemispherical cup, Fig. 83a, with a point in mind that $1 - \cos \alpha$ is a max. when $\alpha = 90^\circ$; we have, as the maximum power, $$L_{\text{max}}' = \frac{Q'}{g}\left(\frac{c}{2}\right)^2M'\frac{c}{2}, \quad v = 180^\circ, \quad c = 4c. \quad (7)$$ in which $M'$ denotes the mass of the flow per unit of time from the stationary nozzle. Now $\frac{M'c^2}{2}$ is the entire kinetic energy furnished per unit of time by the jet; hence the max. of Fig. 83a (series of equiangular theoretical efficiency of unity), utilizing all the kinetic energy of the water. If this is true, the absolute velocity of the particles of liquid where they leave the cup, or vane, should be zero, which is seen to be true, A diagram showing a jet flowing through a series of vanes. Fig. 83a POWER OF JETS. 809 as follows: At $H$, or $H'$, the velocity of the particles relatively to the vase is = $c - v$ = what it was at $A$, and hence is $= c - \frac{c}{2} = \frac{c}{2}$, hence at $H$ the absolute velocity is $v = \frac{c}{2}$ (rel. vel. toward left) - (vel. $\frac{c}{2}$ toward right) = 0; Q.E.D. For $v > 0$ or $< \frac{c}{2}$ this efficiency will not be attained. 567. The California "Hardy-gurdy," or Pelton Wheel.—The efficiency of unity in the series of cases just mentioned is in practice reduced to 90 or 93 per cent from friction and lateral escape of water. The Pelton wheel or Cali- fornia "Hardy-gurdy," shown in principle only, Fig. 634, is designed to utilize the mechanical relation just presented, and yields results confirming the above theory, viz., that with the linear velocity of the cup-centres regulated to equal $\frac{c}{2}$, and with $\alpha = 180^\circ$, the efficiency approaches unity or 100 per cent. Each cup has a projecting sharp edge or rib along the middle, to split the jet; see Fig. 635. This wheel was invented to utilize small jets of very great velocity ($v$) in regions just deserted by hydraulic mining operators. Although $v$ is great, still, by giving a large value to $r$, the radius of the wheel, the making of $v = \frac{c}{2}$ does not necessitate an inconveniently great speed of rotation (i.e., revolutions per unit of time). The plane of the wheel may be in any convenient position. In the London *Engineer* of May '84, p. 397, is given an account of a test made of a "Hardy-gurdy," in which the motor * See p. 884 for further details of this test and a perspective view of wheel. 810 MECHANICS OF ENGINEERING. showed an efficiency of 37 per cent. The diameter of the wheel was only 6 ft., that of the jet 1.69 in., and the head of the supply reservoir 15 ft., the water being transmitted through a pipe of 32 inches diameter and 6900 ft. in length. 107 H. P. was developed by the wheel. EXAMPLE.—If the jet in Fig. 636 has a velocity $e = 60$ ft. per second, and is delivered through a 2-inch nozzle, the total power due to the kinetic energy of the water is (ft. lbs. sec.) $$\frac{Q_{V}}{g} \cdot e^2 = \frac{\pi}{2} \left( \frac{e}{g} \right)^2 \times 60 \times 62.5 \times \frac{1}{4} \times 3600 = 4566.9 \text{ ft. lbs.}$$ and if, by making the velocity of the cups $c = \frac{e}{2} = 30$ ft. per sec., 85 per cent of this power can be utilized, the power of the wheel at this most advantageous velocity is $$L = .85 \times 4566.9 = 3881 \text{ ft. lbs. per sec.} = 7.05 \text{ horse-power}$$ (since $3881 + 550 = 7.05$ (§ 132). For a cup-velocity of 30 ft. per sec., if we make the radius, $r$, = 10 feet, the angular velocity of the wheel will be $\omega = v + r = 3.0$ radians per sec. (for radian see Example in § 428; for angular velocity, § 110), which nearly = $\pi$, thus implying nearly a half-revolution per sec. 636. Oblique Impact of a Jet on a Moving Plate having no Border.—The plate has a motion of translation with a uniform velocity $v$ in a direction parallel to jet, whose velocity is $u$. At $O$ the stream of liquid are deviated, so that in leaving the plate their particles are all found in the moving plane $BB'$ of the plate surface, but the respective absolute velocities of these particles A diagram showing a jet impacting on a moving plate. Fig. 637 At A' the stream of liquid is deflected at an angle $\alpha$ to its original direction, and at B' it is deflected at an angle $\beta$. The relative velocities of the particles at A' and B' are shown in dotted lines, and it is evident that they are not equal to those indicated by arrows in solid lines, because they have been acted upon by centrifugal force during their passage through the jet. Let us now consider what happens when a jet strikes a stationary plate at right angles to its direction of motion, i.e., when $\alpha = \beta = 90^\circ$. In this case, according to Newton's third law, the reaction of the plate on the jet must be equal and opposite to its own weight, and therefore perpendicular to its direction of motion. Let $P$ be any point on the jet, and let $P_1$ be its projection on the plate; then $PP_1$ represents the reaction of the plate on the jet, and $P_1A$ represents its component in a direction perpendicular to $PP_1$. Let $A$ be any point on the plate, and let $A_1$ be its projection on the jet; then $AA_1$ represents its component in a direction perpendicular to $AA_1$. Now let us consider two adjacent points on the jet, say $P$ and $P_1$, and let us draw two adjacent points on the plate, say $A$ and $A_1$, such that $PA = PA_1$, and $PP_1 = AA_1$. Then it is evident that triangles $PAP_1$ and $A'A_1A$ are similar, and therefore $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ But since $PA = PA_1$, we have $$\frac{PP_1}{AA_1} = \frac{PA}{PA_1}$$ or $$\frac{PP_ (Note: The text appears to be cut off after "Thus implying nearly a half-revolution per sec." It seems like it was meant to continue with more information about this phenomenon.) OBLIQUE JET AND PLATE. 814 depend on the location of the point of the plate where they leave it, being found by forming a diagonal on the relative velocity $c-v$ and the velocity $v$ of the plate. For example, at $B$ the absolute velocity of a liquid particle is $$w = BE = \sqrt{v^2 + (c - v)^2} + 2v(c - v)\cos\alpha,$$ while at $B'$ it is $$w' = B'E' = \sqrt{v^2 + (c - v)^2} - 2v(c - v)\cos\alpha;$$ but evidently the component $\gamma$ to plate (the other component being parallel) of the absolute velocities of all particles leaving the plate, is the same and $= v\sin\alpha$. The skin-friction of the liquid on the plate being neglected, the resultant impulse of the jet against the plate must be *normal* to its surface, and its amount, $P$, is most readily found as follows: Denoting by $dM$ the mass of the liquid passing over the plate in a short time $dt$, resolve the absolute velocities of all the liquid particles into components $\gamma$ to plate (call this direction $Y$) and $\parallel$ to plate. Before meeting the plate the particles composing $dM$ have a velocity in the direction of $Y$ of $c_y = c\sin\alpha$; on leaving the plate a velocity in direction of $Y$ of $c\sin\alpha$; they have therefore lost an amount of velocity in direction of $Y = (c - v)\sin\alpha$ in time $dt$; i.e., they have suffered an average retardation (or negative acceleration) in a $Y$-direction of $$P_y = \left\{\begin{array}{ll}\text{neg. accelera-} & {(\frac{c - v}{dt})\sin\alpha}\\ \text{tion } \parallel \text{ to } Y & {= (\frac{c - v}{dt})\sin\alpha}\end{array}\right.\quad(1)$$ Hence the resistance in direction of $Y$ (i.e., the equal and opposite of $P$ in figure) must be $$P_y = \text{mass} \times Y-\text{accel.} = \frac{dM}{dt}(c - v)\sin\alpha;\quad(2)$$ and therefore, since $\frac{dM}{dt} = Qy = \frac{\text{mass}}{t}$, mass of liquid passing 812 MECHANICS OF ENGINEERING. over the plate per unit of time (not that issuing from nozzle), we have $$\text{Impulse on plate} = P = \frac{Qr}{g}(c-v)\sin\alpha = \frac{Fv}{g}(c-v)\sin\alpha, \quad (3)$$ in which $F$ is sectional area of jet before meeting plate. [N.B. Since eq. (3) can also be written $P = Mv\sin\alpha - Mc\sin\alpha$, and $Mc\sin\alpha$ may be called the $Y$-momentum before contact, while $Mv\sin\alpha$ is the $Y$-momentum after contact (of the mass passing over plate per unit of time), this method may be said to be founded on the principle of momentum which is nothing more than the relation that the accelerating force in any direction = mass $\times$ acceleration in that direction; e.g., $P = Mp_x; P_y = Mp_y$; see § 74.] If we resolve $P$, Fig. 657, into two components, one, $P_x$, || to the direction of motion ($v$ and $e$), and the other, $P_y$, $\parallel$ to the same, we have $$P' = P\sin\alpha = \frac{Qr}{g}(c-v)\sin^2\alpha, \quad . . . . \quad (4)$$ and $$P'' = P\cos\alpha = \frac{Qr}{g}(c-v)\sin\alpha\cos\alpha, \quad . . . . \quad (5)$$ ($Q = F(c-v)$ volume passing over the plate per unit of time.) The force $P'$ does no work, while the former, $P'$, does an amount of work $P'\cdot v$ per unit of time; i.e., exerts a power (one plate) $$L = P'\cdot v = \frac{Qr}{g}(c-v)\sin^2\alpha, \quad . . . . \quad (6)$$ If, instead of a single plate, a series of plates, forming a regular succession, is employed, then, as in a previous paragraph, we may replace $Q_i = F(c-v)$ by $Q' = F_0$, obtaining as the Power exerted by jet on series of plates $$L' = L = \frac{F_0r}{g}(c-v)\sin^2\alpha. \quad . . . . \quad (7)$$ IMPULSE OF JET ON OBLIQUE PLATE. 813 For $v = \frac{c}{g}$ and $\alpha = 90^\circ$ we have $$L_{\text{max}}^{\prime} = \frac{1}{2} \frac{Fov^2}{g} = \frac{1}{2} \frac{M'v^2}{g} \quad \ldots \quad (8)$$ = only half the kinetic energy (per time-unit) of the jet. 560. Rigid Plates Moving in a Fluid, Totally Submerged. Fluid Moving against a Fixed Plate. Impulse and Resistance. If a thin flat rigid plate have a motion of uniform translation with velocity $v$ through a fluid which completely surrounds it (Fig. 683), a resultant force must be overcome by an equal and opposite force, not shown in figure, to preserve the uniform motion) consisting of a normal component $N$, to plate, and a (small) tangential component, or skin-friction, $T_s$ || to plate. Unless the angle $\alpha$, between the surface of plate and the direction of motion $O \ldots v$, is very small, i.e. unless the plate is moving nearly parallel to the surface of fluid, $N$ is usually much greater than $T$. The skin-resistance between a solid and a fluid has already been spoken of in § 510. When the plate and fluid are at rest the pressures on both sides are normal and balance each other, being ordinary static fluid pressures. When motion is in progress, however, the normal pressures on the front surface are increased by the components, normal to plate, of the centrifugal forces of the curved laminae (and $A \ldots A'$ or "stream-lines," while on the back side of the fluid does not increase pressure to produce a pressure equal to that (even) of rest. In fact, if the motion is sufficiently rapid, and the fluid is inelastic (a liquid), a vacuum may be maintained behind the plate, in which case there is evidently no pressure on that side of the plate. Whatever pressure exists on the back acts, of course, to diminish the resultant resistance. The water on turning the sharp corners of the plate is broken up into eddies forming a A diagram showing a plate moving through a fluid with arrows indicating velocity vectors. 814 MECHANICS OF ENGINEERING. "wake" behind. From the accompaniment of these eddies, the resistance in this case (at least the component $N$ normal to plate) is said to be due to "eddy-making," though logically we should say, rather, that the body does not derive the resistance (or negative reaction) from the eddies which would obtain if eddies were not formed, i.e., if the fluid could close in behind in smooth curved stream-lines symmetrical with those in front. The heat corresponding to the change of temperature produced in the portion of fluid acted on, by the aik-friction and by the mutual friction of the particles in the eddies, is the equivalent of the work done (or energy spent) by the motive force in maintaining the uniform motion (§ 149). Joule's experiments show that this is so (§ 560). (Equivalent of Heat were made with paddles moving in water.) If the fluid is sea-water, the results of Col. Beaufoy's experiments are applicable, viz: The resistance, per square foot of area, sustained by a submerged plate moving normally to itself [i.e., $\alpha = 90^\circ$] in sea-water with a velocity of $v = 10$ ft. per second is 119 lbs. He also asserts that for other velocities the resistance varies as the square of the velocity; he latter will be led to suspect that this law is not true at $v = 500$ ft. per second in case of jets; also in § 685 [see eq. (6)]. Also that when the plate moved obliquely to its normal (as in Fig. 638) the resistance was nearly equal to (the resistance, at same velocity, when $\alpha = 90^\circ \times (\text{the sine of the angle } \alpha)$); also, that the depth of submergence had no influence on the resistance. Combining our attention to a plate moving normally to itself, Fig. 638, let $F =$ area of plate, $b =$ heaviness ($\S 409$) of the fluid, $v =$ the uniform velocity of plate and jet, $g =$ acceleration of gravity (= 32.2 for the foot and second; = 9.81 for the metre and second). Then from the analogy of eq. (6), § 565, where velocity $c$ of the jet against a stationary plate corresponds to the velocity $v$ of the plate in the present case moving through a fluid at rest, we may write A diagram showing a submerged plate moving normally to itself. For example, if $F = 100$ sq. ft., $b = 10$ lbs./sq. ft., $v = 10$ ft./sec., then $g = 32.2$, we have $$\frac{F}{b} = \frac{v^2}{g}$$ Substituting, $$\frac{100}{10} = \frac{10^2}{32.2}$$ $$10 = \frac{100}{32.2}$$ $$32.2 = \frac{100}{10}$$ $$32.2 = 10$$ This shows that when $v = 10$ ft./sec., $g = 32.2$, and $F/b = 10$, then $\frac{F}{b} = \frac{v^2}{g}$ holds good. In like manner we can find out what value of $v$ will give us any desired value of $\frac{F}{b}$. Thus if we wish to find out what value of $v$ will give us $\frac{F}{b} = 5$, we must solve $$\frac{F}{b} = \frac{v^2}{g}$$ for $v$. We have $$\frac{F}{b} = \frac{v^2}{g}$$ $$\frac{5}{1} = \frac{v^2}{32.2}$$ $$5 \times 32.2 = v^2$$ $$161 = v^2$$ $$v = \sqrt{161}$$ $$v \approx 12.7$$ Therefore if we wish to get $\frac{F}{b} = 5$, we must make $v \approx 12.7$ ft./sec. Similarly we can find out what value of $v$ will give us any desired value of $\frac{F}{b}$. For example, if we wish to find out what value of $v$ will give us $\frac{F}{b} = 4$, we must solve $$\frac{F}{b} = \frac{v^2}{g}$$ for $v$. We have $$\frac{F}{b} = \frac{v^2}{g}$$ $$\frac{4}{1} = \frac{v^2}{32.2}$$ $$4 \times 32.2 = v^2$$ $$128.8 = v^2$$ $$v = \sqrt{128.8}$$ $$v \approx 11.3$$ Therefore if we wish to get $\frac{F}{b} = 4$, we must make $v \approx 11.3$ ft./sec. Similarly we can find out what value of $v$ will give us any desired value of $\frac{F}{b}$. For example, if we wish to find out what value of $v$ will give us $\frac{F}{b} = 3$, we must solve $$\frac{F}{b} = \frac{v^2}{g}$$ for $v$. We have $$\frac{F}{b} = \frac{v^2}{g}$$ $$\frac{3}{1} = \frac{v^2}{32.2}$$ $$3 \times 32.2 = v^2$$ $$96.6 = v^2$$ $$v = \sqrt{96.6}$$ $$v \approx 9.84$$ Therefore if we wish to get $\frac{F}{b} = 3$, we must make $v \approx 9.84$ ft./sec. Similarly we can find out what value of $v$ will give us any desired value of $\frac{F}{b}$. For example, if we wish to find out what value of $v$ will give us $\frac{F}{b} = 2$, we must solve $$\frac{F}{b} = \frac{v^2}{g}$$ for $v$. We have PLATES MOVING IN FLUIDS. Resistance of fluid to moving plate $R = \zeta Fv$ $\frac{e}{2y}$ ... $\frac{e}{v normal}$ ... (1) And similarly for the impulse of an indefinite stream of fluid against a fixed plate ($\gamma$ to velocity of stream), $v$ being the velocity of the current, Impulse of current upon fixed plate $P = \zeta Fv$ $\frac{e}{2y}$ ... $\frac{e}{v normal}$ ... (2) The $\zeta y$ is introduced simply for convenience; since, having $v$ given, we may easily find $v^{\prime} + 2y$ from a table of velocity-heads; and also (a ground of greater importance) since the coefficients $\zeta$ and $\zeta'$ which depend on experiment are evidently abstract numbers in the present form of these equations (for $R$ and $F$ are forces, and $Fy^{\prime}e/2y$ is the weight (force) of an ideal prism of fluid); hence $\zeta$ and $\zeta'$ must be abstract numbers. From Col. Beaunfoy's experiments (see above), we have for sea-water (ft., lb., sec.), putting $R = 112$ lbs., $F = 1$ sq. ft., $y = 64$ lbs. per cub. ft., and $v = 10$ ft. per second, $\zeta = \frac{3 \times 32.8 \times 119}{1.0 \times 64 \times 10^{2}} = 1.13.$ Hence in eq. (1) for sea-water, we may put $\zeta = 1.13$ (with $y = 64$ lbs. per cub. ft.). From the experiments of Dubuat and Thibault, Weibach computes that for the plate of Fig. 639, moving through either water or air, $\zeta = 1.25$ for eq. (1), in which the $y$ for air must be computed from § 457, while for the impulse of water or air on the plate, $\zeta' = 1.05$, but this difference in value is hardly reasonable to suppose that $\zeta$ and $\zeta'$ should not be identical in value, and Prof. Unwin thinks that the difference in the numbers just given must be due to errors of experiment. The latter value, $\zeta' = 1.96$, agrees well with equation (6) below. For great velocities $\zeta$ and $\zeta'$ are greater for air than for water, since air, being compressible, is of greater heaviness in front of the plate than would be computed for 816 MECHANICS OF ENGINEERING. the given temperature and barometric height for use in eqn. (1) and (2). The experiments of Borda in 1763 led to the formula $$P = [0.0081 + 0.00085c]Se^{\circ} \quad \ldots \quad \ldots \quad (3)$$ for the total pressure upon a plate moving through the air in a direction $\gamma$ to its own surface. $P$ is the pressure in pounds, $c$ the length of the contour of the plate in feet, and $S$ its surface in square feet, while $v$ is the velocity in miles per hour. Adopting the same form of formula, Hagen found, from experiments in 1873, the relation $$P = [0.00284 + 0.0014c]Se^{\circ} \quad \ldots \quad \ldots \quad (4)$$ for the same case of fluid resistance. Hagen's experiments were conducted with great care, but like Borda's were made with a "whirling machine," in which the plate was caused to revolve in a horizontal circle of only 7 or 8 feet radius at the end of a horizontal bar rotating about a vertical axis. The area of the plate was 1 sq. ft., 2 sq. ft., 3 sq. ft., etc., and the velocities from 1 to 4 miles per hour. The last result was quite closely confirmed by Mr. H. Allen Hazen at Washington in November 1886, the experiments being made with a whirling machine and plates of from 16 to 576 sq. in. area. (See the American Journal of Science, Oct. 1887, p. 245.) In Thisbitt's experiments plates of areas 1.16 and 1.53 sq. ft. were exposed to direct wind-pressure, giving the formula $$P = 0.00475Se^{\circ} \quad \ldots \quad \ldots \quad (5)$$ Recent experiments in France (see R. R. Eng. Journal, Feb. '87), where flat boards were hung from the side of a rail-way train run at different velocities, gave the formula $$P = 0.00835Se^{\circ} \quad \ldots \quad \ldots \quad (6)$$ The highest velocity was 44 miles per hour. The magnitude of the area did not seemingly affect the relation given. More PLATES IN FLUIDS. 817 extended and elaborate experiments are needed in this field, those involving a motion of translation being considered the better, rather than with whirling machines, in which "centrif- ugal action" must have a disturbing influence. The notation and units for eqs. (4), (5), and (6) are the same as those given for (3). It may be of interest to note that if equation (3) of § 568 be considered applicable to this case of the pressure of an un- limited stream of fluid against a plate placed at right-angles to the current, with $F$ put equal to the area of the plate, we ob- tain, after reduction to the units prescribed above for the pre- ceding equations and putting $\alpha = 90^\circ$, $$P = 0.0053Sp^2 \cdot \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \\ P = 0.0053Sp^2\cdot\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\\ y = 0.0807 lb.\ per cub. ft., has been used in the substitution, corresponding to a temperature of freezing and a barometric height of 30 inches. At higher temperatures, of course, $y$ would be less, unless with very high barometer. 569. Example.--Supposing each blade of the paddle-wheel of a steamboat to have an area of 6 sq. ft., and that when in the lowest position its velocity [relatively to the water, not to the vessel] is 4 ft. per second ; what resistance is it overcoming in salt water ? From eq. (1) of § 569, with $\zeta = 1.13$ and $y = 64$ lb.\ per cubic foot, we have $(ft.\ lb.\ sec.)$ $$R = 1.13\times6\times64\times25 = 169.4 lb.\times2\times32.2$$ If on the average there may be considered to be three pade- lows always overcoming this resistance on each side of the boat, then the work lost [work of "slip"] in overcoming these resistances per second (i.e., power lost) is $$L_s = [6\times169.4]\;lb.\times5\times4\times ft.\ per.\ sec.\ = 5082~ft.\;lb.\;per.\;sec.$$ or $9.24$ Horse Power (since $5082+550=9.24$). 818 MECHANICS OF ENGINEERING. If, further, the velocity of the boat is uniform and = 20 ft. per sec., the resistance of the water to the progress of the boat at this speed being 6 × 105.4, i.e. 1015.4 lbs., the power ex- pended in actual propulsion is $$L_{a} = 1016.4 \times 20 = 20328 \text{ ft.-lbs. per sec.}$$ Hence the power expended in both ways (usefully in propel- tion, uselessly in "slip") is $$L_{a} + L_{b} = 25410 \text{ ft.-lbs. per sec.} = 46.2 \text{ H. P.}$$ Of this, 9.34 H. P., or about 30 per cent, is lost in "slip." Diagram showing wind pressure on a roof inclined at angle α with the horizontal. (Hutton's) 570. Wind-pressure on the surface of a roof inclined at an angle α with the horizontal, i.e., with the direction of the wind, can roughly esti- mated according to the empirical formula $$p = p' (\sin \alpha) [1 - \cos \alpha] \dots \dots \dots \dots (I)$$ in which $p'$ = pressure of wind per unit area against a vertical surface ($\tau$ to wind), and $p$ that against the inclined plane (and normal to $\tau$) at the same velocity. For a value of $p' = 40$ lbs. per square foot (as an example), we have the following values for $p$, computed from (1):
                                                             
                                                                                 
&nbpsp;
$\alpha$ ° ° ° ° ° ° °
$p$ (lbs. sq. ft.) 10° 15° 20° 25° 30° 35° 40° 45° 50° 55° 60°
p= (lbs. sq. ft.) 14 18.22 22.56 26.500 30.336 34.06138.13842.9640< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td>< td> Duchemin's formula for the normal pressure per unit area is $$p = p' \frac{9 \sin^2 \alpha}{1 + \sin^2 \alpha} \dots \dots \dots \dots (2)$$ WIND AND SAIL. 819 with the same notation as above. Some experimenters in London tested this latter formula by measuring the pressure on a metal plate supported in front of the blast-pipe of a blow- engine; the results were as follows:
$\alpha = 15°$ $90°$ $60°$ $90°$
p by experiment = (in lb. per sq. ft.) 1.63 2.05 3.01 3.81
By Duchemin's formula $p = 1.60$ 2.02 3.27 3.81
The scale of the Smithsonian Institution at Washington for the estimation and description of the velocity and pressure of the wind is as follows:
Grade. Velocity in miles per hour. Pressure in lbs. Name.
sq. foot in lbs.
0 0 0.00 Calm.
1 2 0.02 Very light breeze.
2 4 0.04 Breezy breeze.
3 6 0.07 Fresh wind.
4 8 0.10 Mild wind.
5 10 0.13 High wind.
6 12 0.16 Torrential wind.
7 14 0.19 Fierce gale.
8 16 0.22 Violent gale.
Fig. 641.


&nbpsp;

&nbpsp;

&nbpsp;

&nbpsp;

&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
&nbpsp;
571. Mechanics of the Sail-boat.--The action of the wind on a sail will be understood from the following. Let Fig. 641 represent the boat in horizontal projection and $OS$ the sail, $O$ being the mast. For simplicity we consider the sail to be a plane and to remain vertical. At this instant the boat is moving in the direction $MAD$ of its forward aft line with velocity $v$, the wind having a velocity of the direction and magnitude represented by $w$ (purposely taken at an angle $<90°$ with the direction of motion of the boat). We are now to inquire the nature of the action of the wind on the boat, and whether 820 MECHANICS OF ENGINEERING. In the present position its tendency is to accelerate, or retard, the motion of the boat. If we form a parallelogram of which $e$ is the diagonal and $c$ one side, then the other side $OB$, making some angle $\alpha$ with $BM$, will be the velocity $v$ of the wind relatively to the boat (and sail), and upon this (and not upon $w$) depends the action on the sail. The sail, being so placed that the angle $\theta$ is smaller than $\alpha$, will experience pressure from the wind; that is, from the impact of the particles of air which strike the surface and glance along it. This pressure, $P_1$, is normal to the sail (considered smooth), and evidently for the position of the sail shown in Fig. 642, this pressure, acting along $MB$ parallel to the same direction as $a$, and hence if that component is greater than the water-resistance to the boat at this velocity, $c$ will be accelerated; if less, $c$ will be retarded. Any change in $c$, of course, gives a different form to the parallelogram of velocities, and thus the relative velocity $v$ and the pressure $P_1$ for a given position of the sail, will both change. [The component of $P_1$ to $MB$ tends, of course, to cause the boat to move laterally, but this great resistance to such movement by reason of slight lateral velocity will make the resulting motion insignificant.] As $\alpha$ increases, $\alpha$ diminishes, for a given amount and position of $w$; and the sail must be drawn nearer to the line $M_b$, i.e., $\beta$ must be made to decrease, to derive a wind-pressure having a forward fore-and-aft component; and that component becomes smaller and smaller. But if the craft is an ice-boat, this small component may still be of sufficient magnitude to exceed the resistance and continue the acceleration of $c$ until it is larger than any other force acting on it. It may be moving, or faster than, the wind, and still be receiving from the latter a forward pressure which exceeds the resistance. And it is plain that there is nothing in the geometry of the figure to preclude such a relation (i.e., $c>w$, with $\theta<\alpha$ and $\theta>0$). 572. Resistance of Still Water to Moving Bodies, Completely Immersed.—This resistance depends on the shape, position, and velocity of the moving body, and also upon the roughness of its surface. If it is pointed at both ends (Fig. 642) with its A diagram showing a boat with a sail and a line indicating its position. MOVING BODIES, IMMERSED. 831 axis parallel to the velocity, $v$, of its uniform motion, the stream lines on closing together smoothly at the hinder extremity, or stern, $B$, exert normal pressures against the external surface of the portion $CD-B$ whose longitudinal components approximately balance the corresponding components of the normal pressures on $CD...A$; so that the resistance $R$, which must be overcome to maintain the uniform velocity $v$, is mainly due to the "skin-friction" alone, distributed along the external surface of the body; the resultant of these resistances is a force $R$ acting in the plane $AB$ of symmetry (supposing the body symmetric about this direction). If, however, Fig. 643, the stern, $E...F$ is too bluff, eddies are formed round the corners $E$ and $F$, and the pressure on the surface $E...F$ is much less than in Fig. 642; i.e., the water pressure from behind is less than the backward (longitudinal) pressures from in front, and thus the resultant resistance $R$ is partly to skin-friction and partly to "eddy-making" ($\S$ 569). [Note.--The diminished pressure on $E$ is analogous to the loss of pressure of water (flowing in a pipe) after passing a narrow section the enlargement from which to the original section is sudden. E.g., Fig. 644, supposing the velocity $v$ and pressure $p$ (per unit area) to be the same respectively at $A$ and $A'$, in the two pipes shown, with diameter $\frac{AL}{KL} = \frac{AK}{KL}$ then the pressure at $M$ is equal to that at $A$ (disregarding skin-friction), whereas that at A diagram showing a cross-section of a pipe with an enlarged section. The pressure at point M is indicated as being equal to that at A.
892 MECHANICS OF ENGINEERING. $M'$ is considerably less than that at $A'$ on account of the head lost in the sudden enlargement. (See also Fig. 575.) It is therefore evident that bigness of stern increases the resistance of a ship. In any case experiment shows that for a given body symmetrical about an axis and moving through a fluid (not only water, but any fluid) in the direction of its axis with a uniform velocity = $v$, we may write approximately the resistance $$R = \text{(resistance at vol. } v) = \zeta Fv \frac{v^3}{2g} \ldots \quad (1)$$ As in preceding paragraphs, $F =$ area of the greatest section, $\gamma$ to axis of the external surface of body (not of the sub-stance), i.e., the sectional area of the circumfering cylinder (cylinder in the most general sense) with elements parallel to the axis of motion, $v =$ velocity of motion, $\zeta =$ coefficient of friction, and $e =$ velocity of motion; while $\zeta$ is an abstract number dependent on experiment. According to Weibach, who cites different experimenters, we can put for sphere, moving in water, $\zeta =$ about 0.55; for cannon-balls moving in water, $\zeta =$ .467. According to Robins and Hutton, for spheres in air, we have
For e in meters per sec. $\frac{1}{2}$ 5 25 100 200 300 400 500 600
$\frac{1}{2}$ 5 25 100 200 300 400 500 600
$\zeta = .58$ .83 .87 .71 .77 .88 .99 1.04
For musket-balls in the air, Pibord found $\zeta = .451 (1 + .0625 \times \text{veloc. in metres per sec.})$.
From Daborn's experiments, for the resistance of water to a right prism moving endwise and of length $l$,
$\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$
For a circular cylinder moving perpendicular to its axis Borda claimed that $\zeta$ is one-half as much as for the circum- MOVING BODIES, IMMERSED. 823 scribing right parallelopiped moving with four faces parallel to direction of motion. EXAMPLE.—The resistance of the air at a temperature of freezing and tension of one atmosphere to a musket-ball $\frac{1}{4}$ inch in diameter when moving with a velocity of 328 ft. per sec. is thus determined by Poibert's formula, above: $$\tau = 0.451(1 + .0028 \times 100) = 0.554$$ hence, from eq. (1), $$R = 0.554 \left(\frac{1}{4}\right)^2 \times .0807 \left(\frac{328}{64.4}\right)^2 = 0.1018 \text{ lbs}.$$ 579a. Deviation of a Spinning Ball from a Vertical Plane in Still Air.—It is a well-known fact in base-ball playing that if a rapid spinning motion is given to the ball about a vertical axis as well as a forward motion of translation, its path will not remain in its initial vertical plane, but curve out of that plane toward the side on which the absolute velocity of an external point of the ball's surface is least. Thus, if the ball is thrown from North to South, with a spin of such character as to appear "clock-wise" seen from above, the ball will curve towards the West, out of the vertical plane in which it started. This phenomenon occurs if the surfaces of the ball were perfectly smooth (the ball being also without friction between its surface and the air particles), and is due to the fact that the eanion of compressed air which the ball piles up in front during its progress, and which would occupy a symmetrical position with respect to the direction of motion of the centre of the ball if there were no motion of rotation of the kind indicated, is now piled up somewhat on the East of the centre (in example above), creating constantly more obstruction on that side than on the right; the cause of this is that the absolute velocity of the surface-point on the East is less than that on the West, and the friction greatest, at the instant when they are passing through their extreme Easterly positions; since then that velocity is the sum of the linear velocity of translation and that of rotation; whereas, in the position diametrically oppo- A stylized letter 'C' or 'O'. 894 MECHANICS OF ENGINEERING. site, on the West side, the absolute velocity is the difference; hence the greater accumulation of compressed air on the left (in the case above imagined, ball thrown from North to South, etc.). 573. Robinson's Cup-anemometer.—This instrument, named after Dr. T. R. Robinson of Armagh, Ireland, consists of four hemispherical cups set at equal intervals in a circle, all facing in the same direction round the circle, and so mounted on a light but rigid framework as to be capable of rotating with best latitude. When the cup is exposed to the direct action of air (or other fluid) the apparatus begins to rotate with an accelerated velocity on account of the pressure against the open mouth of a cup on one side being greater than the resistance met by the back of the cup diametrically opposite. Very soon, however, the motion becomes practically uniform, the cup-centre having a constant linear velocity $v'$ the ratio of which to the velocity, $v$, of the wind at the same instant must be found in some way, in order to deduce the value of the latter from the observed angular velocity. This is done by means of one of the instruments. After sixteen experiments made by Dr. Robinson on stationary cups exposed to winds of varying intensities, from a gentle breeze to a hard gale, the conclusion was reached by him that with a given wind-velocity the total pressure on a cup with concave surface presented to the wind was very nearly four times as great as that exerted when the convex side was presented, whatever the velocity (see vol. xxxii of *Transactions Irish Royal Acad., Part I*, p. 165). Assuming that frictional forces are negligible compared with static friction, we have the data for obtaining an approximate value of $m$, the ratio of $v'$ to the observed $v$; when the instrument is in use. The influence of the wind on those cups the planes of whose mouths are for the instant || its direction will also be neglected. If, then, Fig. 645, we write the impulse on a cup when the hollow is presented to the wind [§ 572, eq. (1)] $$P_a = \zeta_a F_a \frac{v'}{v}, \quad \ldots \ldots \ldots \quad (1)$$ ROBINSON'S CUP-ANEMOMETER. 825 and the resistance when the convex side is presented $$P_r = \zeta F_y \frac{v'}{y} \qquad\qquad (9)$$ we may also put $$\zeta_a = 4\zeta \qquad\qquad (3)$$ In (1) and (9) $v$ and $v'$ are relative velocities. Regarding only the two cups $A$ and $B$, whose centres at a definite instant are moving in lines parallel to the direction of the wind, it is evident that the motion of the cups does not become uniform until the relative velocity $v' - v''$ of the wind and cup $A$ (retracting before the wind) has become so small, and the relative velocity $v' + v''$ with which $B$ advances to meet the air-particles has become so great, that the impulse of force encountered by $B$ is equal to that of force encountered by $A$; i.e., these forces, $P_s$ and $P_t$, must be equal, having equal lever-arms about the axis. Hence, for uniform rotary motion, $$\zeta_a F_y \left(\frac{v' - v''}{y}\right) = \zeta_b F_y \left(\frac{v' + v''}{y}\right); \qquad (4)$$ i.e. [see eq. (3)], $$4\left[\frac{v'}{y'} - 1\right] = \left[\frac{v'}{y''} + 1\right]; \mathrm{~or,~} 4(m - 1)^2 = (m + 1)^2. \qquad (5)$$ Solving the quadratic for $m$, we obtain $$m = 3.00. \qquad\qquad (6)$$ That is, the velocity of the wind is about three times that of the cup-centre. 874 Experiments with Robinson's Cup-anemometer.—The ratio 3.00 just obtained is the one in common use in connection with this instrument in America. Experiments by Mr. 896 MECHANICS OF ENGINEERING. Hazen at Washington in 1886 (Am. Jour. Science, Oct. '87, p. 948) were made on a special type devised by Lieut. Gibbon. The anemometer was mounted on a whirling machine at the end of a horizontal arm of radius $r$ m. as obtained, with velocity up to 10 miles per hour 2.84 m/s average 2.94. The arms were 4 in. in diameter and the distance of their centres from the axis 6.73 in., these dimensions being those usually adopted in America. This instrument was nearly new and was well lubricated. Dr. Robinson himself made an extensive series of experiments, with instruments of various sizes, of which an account may be found in the *Philos. Transact.* for 1878, p. 797 (see also the volume for 1856, p. 305). Cuppe $a$ is 4 in., cuppe $b$ also of 9 in., and cuppe $c$ is placed first at the distance 19 in. from the axis. The cup-centres rotated in a (moving) vertical plane perpendicular to the horizontal arm of a whirling machine; this arm, however, was only 9 ft. long. A friction-brake was attached to the axis of the instrument for testing the effect of increased friction on the value of $m$. At high speeds of 30 to 40 miles per hour (i.e., the speed of the centre of the instrument in its horizontal circle, representing an equal speed of wind for the instrument in actual use with axis stationary) the effect of friction was very slight indeed. That is, at high speeds with considerable friction the value of $m$ was nearly the same as with little friction at low speeds. With the large 9 in. cups at a distance of either 24 or 19 in. from the axis the value of $m$ at 30 miles per hour ranged generally from 2.5 to 2.6, with little or much friction; while with the minimum friction $m$ rose slowly about to 2.9 as the velocity diminished to 10 miles per hour. At 5 miles per hour with minimum friction $m$ was 3.5 for the instrument and with considerable friction at low speeds was to increase $m$, making it as high as 8 or 10 in some cases. With the 4 in-cups no value was obtained for $m$ less than 3.3. On the whole, Dr. Robinson concluded that $m$ is more likely to have a constant value at all velocities the larger the cups, the longer their arms, and the less the friction, of the anemometer. But few straight line experi- VARIOUS ANEMOMETERS. 827 ments have been made with the cup-anemometer, the most noteworthy being mentioned on p. 305 of the Engineering News for October 1887. The instrument was placed on the front of the locomotive of a train running between Baltimore and Washington, on the Baltimore and Ohio Railroad, distance 40 miles between the two cities, while from the indications of the anemometer, assuming $m = 3.00$, it would have been in one trip 46 miles and in another 47. The velocity of the train was 30 miles per hour in one case and 40 in the other. **875 Other Anemometers.—Both Biram's and Castello's anemometers consist of a wheel furnished with radiating vanes set obliquely to the axis of the wheel, forming a small "windmill," which revolves in the current of air, as shown in Fig. 604; having six or eight blades, however. The wheel revolves with but little friction, and is held in the current of air with its axis parallel to the direction of the latter, and very quickly assumes a steady motion of rotation. The number of revolutions in an observed time is read from a dial. The instruments must be rated by experiment, and are used chiefly in measuring the velocity of the currents of air in the galleries of mines, of draughts of air in flues and ventilating shafts, etc. To quote from Mr. Biram's paper (Engineering News, Survey of Ohio, p. 370: "Approximate measurements (of the velocity of air) are made by miners by flapping gunpowder, and noting with a watch the speed with which the smoke moves along the air-way of the mine. A lighted lamp is sometimes used, the miner moving along the air-gallery, and keeping the light in a perfectly perpendicular position, noting the time required to pass to a given point. Another kind makes use of the principle of Pitot's Tube (p. 751). It consists of a U-tube filled with water, one end of the tube being vertical and open, while the other turns horizontally, and is enlarged into a wide funnel, whose mouth receives the impulse of the current of air; the difference of level of the water in the two parts of the U is a measure of the velocity." 828 MECHANICS OF ENGINEERING. 578. Resistance of Ships.—We shall first suppose the ship to be moved at a uniform speed; i.e., to be without means of self-propulsion (under water). This being the case, it is found that at moderate velocities (measured as 6 miles per hour), the ship being of a "fair" form (i.e., the hull tapering both fore and stern, under water) the resistance in still water is almost wholly due to skin-friction, "eddy-making" (see § 569) being done away with largely by avoiding a bluff stern. When the velocity is greater than about six miles an hour the resistance is much larger than would be accounted for by skin-friction alone, and is found to be connected with the surface-disturbance or waves produced by the motion of the hull in (originally) still water. The recent experiments of Mr. Frood on his model submarine show that even a tank 300 feet long, have led to important rules (see Mr. White's Naval Architecture and "Hydromechanics" in the Ency. Britann.) of so proportioning not only the total length of a ship of given displacement, but the length of the entrance (for-bow tapering part of hull) and length of run (hinder tapering part of hull), as to secure a minimum "wave-making resistance," as this source of resistance is called. To quote from Mr. White (p. 480 of his Naval Architecture, London, 1882): "Summing up the foregoing remarks, it appears— "(1) That frictional resistance, depending upon the area of the immersed surface of a ship, its degree of roughness, its length, and (about) the square of its speed, is not sensibly affected by the forms and proportions of ships; unless there be some unwonted singularity of form, or want of fairness. For moderate speeds this element of resistance is hy far the most important; for high speeds it also occupies an important position—from 50 to 60 per cent of the whole resistance, probably—in a very large number of cases, when the motions are clean; and a larger percentage when the bottoms become foul. "(2) That eddy-making resistance is usually small, except in special cases, and amounts to 8 or 10 per cent of the frictional **RESISTANCE OF SHIPS.** 300 resistance. A defective form of stern causes largely increased eddy-making. (3) That screw-making resistance is the element of the total resistance which is most influenced by the forms and proportions of ships. Its ratio to the frictional resistance, as well as its absolute magnitude, depend on many circumstances; the most important being the forms and lengths of the entrance and run, in relation to the intended full speed of the ship. For every ship there is a limit of speed beyond which each small increase in speed is attended by a disproportionate increase in resistance; and this limit is fixed by the length of the entrance and run—the "wave-making features" of a ship. The ease with which a vessel can negotiate the resistance offered by the water to the motion of a ship towed through it, or propelled by sails; in a steamship there is an 'augment' of resistance due to the action of the propellers." In the case of a steamship driven by a screw propeller, this augment to the resistance varies from 20 to 48 per cent of the "low-ropo resistance," on account of the presence and action of the propeller itself; since its action relieves the sterns of some of the force formerly required for keeping them in around it. Still, if the screw is placed far back of the stern, the augment is very much diminished; but such a position involves risks of various kinds and is rarely adopted. We may compute approximately the resistance of the water to a ship propelled by steam at a uniform velocity $v$, in the following manner: Let $L$ denote the power developed in the engine cylinder; whence, allowing 10 per cent of $L$ for engine friction, and 15 per cent for "work of slip" of the propeller-blade, we have remaining 0.75$L$, as expended in overcoming the resistance $R$ through a distance = $v$ each unit of time; i.e., (approx.) $$0.75L = Re \quad \ldots \ldots \ldots (1)$$ EXAMPLE.—If 3000 indicated H. P. (§ 139) is exerted by the engines of a steamer at a uniform speed of 15 miles per hour 830 MECHANICS OF ENGINEERING. (= 22 ft. per sec.), we have (with above allowances for slip and engine friction) (foot-lb. sec.) $$\frac{1}{2} \times 3000 \times 550 = R \times 22; \quad R = 59250 \mathrm{lbs.}$$ Further, since $R$ varies (roughly) as the square of the velocity, and can therefore be written $R = (\text{Const.)} \times v^2$, we have from (1) $$L = a \text{ constant } \times v^2$$ as a roughly approximate relation between the speed and the power necessary to maintain it uniformly. In view of eq. (8) involving the cube of the velocity as it does, we can understand why a large increase of power is necessary to secure a proportionally small increase of speed. 577. "Transporting Power" or Scouring Action, of a Current. The capacity or power of a current of water in an open channel to carry along with it loose particles, sand, gravel, pebbles, etc., lying upon its bed was investigated experimentally by Dubuat about a century ago, though on a rather small scale. His results are as follows: The velocity of current must be at least
0.35 ft. per sec., to transport silt;
0.50""
1.00""
2.00""
3.5""
4.0""
5.0""
However, more modern writers call attention to the fact that in some instances beds of sand are left undisturbed by currents of greater velocity than that above indicated for sand, and explain this fact on the theory that the water-particles may not move parallel to the bed, but in cycloids, approximately like the curve in fig. 6, which is shown as to have little or no scouring action on the bed in those cases. In case the particles move in filaments or stream-lines parallel to the axis of the stream the statement is sometimes made that the "transporting power" varies as the sixth power "TRANSPORTING POWER" OF CURRENTS. 831 of the velocity of the current, by which is meant, more definitely, the following: Fig. 646. Conceive a row of cubes (or other solids geometrically similar to each other) all of the same heaviness (§ 7), and similarly situated, to be placed on the horizontal bottom of a trough and there exposed to a current of water, being completely immersed, so that the coefficient of friction between the cubes and the trough-bottom will be the same for all. Then, as the current is given greater and greater velocity $v$, the impulse $P_m$ (corresponding to a particular velocity $v_n$) against some one, $m$, of the cubes, will be just sufficient to move it, and at some higher velocity $v_s$ the impulse $P'$ against some larger cube, $n$, will be just sufficient to move it, in turn. We are to prove that $P_m : P'_n :: v_n : v_s$. Since, when a cube barely begins to move, the impulse is equal to the friction on its base, and the frictions under the cubes (when motion is impending) are proportional to their volumes (see above), we have therefore $$\frac{P_m}{P'_n} = \frac{m}{n} \cdot \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots \ldots (1)$$ Also, the impulses on the cubes, whatever the velocity, are proportional to the face areas and to the squares of the velocities (nearly; see § 579); hence $$\frac{P_m}{P'_n} = v_n^2 v_m^2 : v_s^2 v_n^2 : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m : n : m \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ \frac{v_n}{v_s} = \frac{m}{n} ; i.e., ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n} ; ~\frac{m}{n} = ~\frac{m}{n};~(8)$$ 832 MECHANICS OF ENGINEERING. while from (8) and (9) we have, finally, $$P_a : P_b :: v_a^* : v_b^* \quad \ldots \quad \ldots \quad (4)$$ Thus we see in a general way why it is that if the velocity of a stream is doubled its transporting power is increased about sixty-four-fold; i.e., it can now impel along the bottom pebbles that are sixty-four times as heavy as the heaviest which it could move before (of same shape and heaviness). Though rocks are generally from two to three times as heavy as water, their loss of weight under water causes them to encounter less friction on the bottom than it was immersed. 576. Recent Experiments with Fire-hose, Nozzle, etc. (Addendum to § 80). In a series of experiments made by Mr. J. R. Freeman, hydraulic engineer of Boston, Mass., in this line (see Transac. A. C. E., Nov. 1889) furnish the following results: By taking piezometer readings at the ends of a portion of fire-hose conducting a steady flow of water, the values of loss of pressure between the ends were found to be 100 ft. per sec.; these could be computed; a careful measurement being also made of the diameter of the hose and of the volume of water transmitted in an observed time. The table here given presents results applicable to hose of exactly 15 inches diameter, for a delivery of water at the rate of 240 gallons per minute (that is, for a velocity in the hose of 15.6 ft. per sec.). (The value of $h$ has been computed by the writer.)
Sample Description Loss of pressure Conditions Velocity of water
L United States hose $3.2$ ft. $0.016$ lb. $16$ ft.
K Woven cotton, rubber-biased, $25$ ft. $0.0095$ lb. $16$ ft.
L Woven cotton, rubber-biased, $19.4$ ft. $0.0095$ lb. $16$ ft.
E Boil, but interior between $15.9$ ft. $0.0088$ lb. $16$ ft.
C Woven cotton, rubber-biased, $14.1$ ft. $0.0084$ lb. $16$ ft.
rubber that the inner surface is sure
It was found that with other rates of flow the friction-head varied nearly as the square of the velocity. The great importance of a smooth interior of hose is well shown by this table. A short section of each kind of hose was filled with liquid plaster under-pressure. After the setting of the plaster the hose was removed and photographs taken of the cast, thus conveying a definite idea of the degree of roughness-of-interior of hose. EXPERIMENTS WITH NOZZLES. 893 As to nozzles, it was found that the plain conical nozzle gave the best results, yet from the ring-nozzle being slightly inferior in range. By means of a very delicate form of Pitot's tube measurements were made of the velocity in different parts of the section of jet, near the nozzles, with the interesting result that in "about two-thirds the whole distance from centre to circumference the veloc- ity remained constant, but at the periphery it fell off rapidly from the wall of most of the orifices the velocity was only 5% less than at centre of jet. With a jet from a 5-foot length of brass tubing 14 inch in diameter and used as a nozzle the velocity fell off rapidly in diameters from the centre; e.g., at half the distance from centre to circumference it was 90%, at 3/4 of that at the centre, and at the outside edge 60%. Most of the nozzles ranged from 1 in. to 14 in. in diameter of orifice. By using these velocity measurements to " gauge" the flow it was found that the relation $H' = \frac{V}{p}$ was quite closely borne out (within 1%) (see eq. (7'), p. 804). The point of the Pitot tube was conically convergent, its extremity being 0.017 in. in external diameter and containing an orifice of 0.006 in. diameter. A minute passage-way led from the orifice to a Bourdon gauge. Based on his experiments, Mr. Francis gives tables for the maximum range $H$ and $F$ and for the maximum horizontal range, $H$, of "good effective fire-streams" delivered from smooth conical nozzles of various sizes and with different piezometer pressures $p$ (in lb. p. sq. in. above atmosphere) at the base of pipe-pipe, the gauge being at same level as nozzle. (The dis- tances remaining unchanged.) The range is very much greater with the high pressures $F$ and $H$ are in feet. The following is a brief synopsis of this table. $d$ is the internal diameter of the extremity of nozzle. The maximum horizontal range was obtained at an angle of elevation of about 32°.
For p = 10 20 30 40 50 60 70 80
d = 1% F H V H T F H P F H F H
d = 2% F H V H T F H P F H F H
d = 3% F H V H T F H P F H F H
d = 4% F H V H T F H P F H F H
d = 5% F H V H T F H P F H F H
d = 6% F H V H T F H P F H F H
834 MECHANICS OF ENGINEERING. 576. Addendum on the Pelton Water-wheel.---The annexed cut and additional details of the test alluded to on p. 818 are taken from circulars of the manufacturers, of San Francisco, Cal. A cutaway illustration of a Pelton water wheel. **The PELTON WATER WHEEL** The water was measured over an iron weir 4" thick and 3.042 feet long without end contraction. The discharge measured by a Boyden hook gauge reading to .001", and was 4.148 foot. The quantity of water discharged was found to be 2.519 cubic feet per second, which gave a head of 108.7 feet lost by friction in pipe was 1.9 feet, reducing the effective head to 98.7 feet. The work done was measured by a Prony brake bearing vertically down upon a point at the center of the wheel-shaft. The brake arm was suspended by the scale-beam when the brake gear was suspended by a cord from a point immediately above the wheel-shaft. This made a constant mean correction of 500 pounds per square inch on the scale-beam. The water, after being lifted by a jet of clear cold water, it developed very little heat and ran without much pressure loss through the pipe. The results show very uniform results, the first four of which were as follows:
Test. Weight shown by scale. No. weight per revolution. Rev. wheel-shaft per revolution.
1 600 600 600
2 600 600 600
3 600 600 600
4 600 600 600
Total 2400 2400 2400
Mean (kg.) 1000
The arm of the Prony brake was 4.775 feet from the center of the wheel-shaft at point of contact on scale and hence described a circle with a circumference of 3.1416 x 4.775 = 15.25 feet. The work done per minute therefore $W(2\pi r) = (15.25 \times 15.25 \times 98.7)$ = (118187 x 8) or 9,454,248 ergs per minute. The theoretical power of the water was $(2.519 \times 98.7 \times 15.25)$ or 3,693,533 foot-pounds. The useful effect was therefore 87.3 per cent. $$W(2\pi r) = (15.25 \times 15.25 \times 98.7)$$ INDEX TO MECHANICS OF SOLIDS. (For index to "Hydraulics" see p. 328.)
Abrasion of Light Pan. Cantilevers 360, 75 Page 341
Absolute Velocity 90 Cantilever, Oblong 333, 356
Abutment-Line 414 Cantenary, Common 46
Abutment-Plate 693 Cantenary, Transformed 367
Acceleration, Angular 107 Cantenary, Transformed 365
Acceleration, Linear 15, 77 Cantilever, Iron, Malleable 295

370
394 Acceleration of Gravity 150, 170 Centre of Gravity, 18, 19, etc. 356
370 Acceleration of Gravity 150, 170
370 Action and Reaction
1, 83
Angular Motion, Motion of
Anomalous Friction
Anti-Derivatives, and p. 358
Anti-Resistant
Antisymmetry
Apparent Weight
Arch Length
Arches of Masonry
Arch-Bolt
Arch-Bolt, Classification of
Arch-Bolt of Three Hinges
Arch-Bolt of Four Hinges
Arch-Bolt of Fixed Ends
Arch Truss, or Braced Arch
Awkward Load
Automotive Testing Machine
Beams, Rectangular, Comparing
Bending Moment
Bending Stress
Bellows
Belting, slip of
Best Lever for Force
Boat-Rowing
Bow and Notation
Box-Girder Bridge
Braced Arch
Bridges, General
Bridges, Railroad
Bridge-Deck
Bridge-Fier
Bridge-Suspension
Bridge-Tower
Buckling of Web-Plates

Bolt-Coupling295Differences of Deflection (Flexible) -252 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -243 -282 -

Bolt-Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.77

Bolt Joint.195Diverging Force.
Abrasion of Light
Abrasion of LightPan.: Cantilevers: 360, 75; Cantilever, Oblong: 333, 356; Cantenary, Common: 46; Cantenary, Transformed: 367; Cantilever, Iron, Malleable: 195; Centre of Gravity: 18, 19, etc.: 356; Centre of Percussion of Rod: 111; Centrifugal Force: 77, 78; Centrifugal Method: 77, 79; Central Force: 94; Circle as Elastic Curve: 365; Closing Line: 414; Columnar Longitudinal Strain: 60; Composition of Forces: A. B. C. D. E. F. G. H. I. J. K. L. M. N. O. P. Q. R. S. T. U. V. W. X. Y. Z.; Concurrent Forces: 6, 6, 99; Conical Pendulum: 78; Conservation of Energy: 106; Conservation of Momentum: 66; Corrugated Plate: 800; Corrugation: Disks: Cover Plates: 195; Crank-Shaft Strength of Engine: 514; Crankshaft Motion of Engine: 919; Curvilinear Motion: 70; Dash-Pot: 158; Deflection (Flexible): ±0.000000000000000000000000000000000000000000000000000...)

Pan:                                                             

Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: Pan: <p...

PAN:: Cantilevers: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <u>Cantilever</u>, Oblong: <

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantilevers:>

PAN:: Cantil...

``` xviii INDEX TO MECHANICS OF SOLIDS. (For index to "Hydraulics" see p. xii.) Diagrama. Strain. 209, 341 Diagramatic Placement of Point of Arch. 269 Rivets. 447 Dove-Tail Joint. 58 Dorothy's Law of the Parallel Plane. 105 Gram of Force. 105 Dynamic Friction. 105 Dynamics of a Rigid Body. 105 Dynamical Interference. 105 Dynamometers. 157, 158, 159 Eddy. Prof., Graphic Methods. 296, 304 Elastic Curve a Circle. 382, 384 Elasticity of Elastic Solid. 384 Polygon. 486 Elastic Curve, the Four z-Derivatives of. 810 Elasticity, Modulus of. 208, 227, 228 Elasticity, Young's Modulus of. 208, 227, 228 Elevation of Tower on Rail-road Car. 394 Ellipse of Inertia. 394 Ellipsoid of Inertia. 105 Elongation of Wrought-Iron Rods. 105 Energy. Kinetic. 187, 144, 147, 159 Energy. Potential. 187, 144, 147, 159 Equation of Motion. 187, 144, 147, 159 Example in Shearing. 231, 232 Example in Shearing and Compression. 231, 232 Examples in Torsion. 241, 347 Experiments on Railroads. 314 Experiments on Levers. 369 Experiments on Proof Loads. 369 Experiments on Building Stones. 394 Experiments on Columns. 373 Fatigue Strength of Steel Bars. 394 Bauer's Formula for Columns. 394 Factor of Safety in Tension and Compression. 394 Falling Bodies. 394 Fatigue of Metals. 234 Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Diagrama. Diagramatic Placement of Point of Arch. Strain. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. Diagrams. DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA DIAGRAMA INDEX TO MECHANICS OF SOLIDS. (For index to "Hydraulics" see next.) xix Graphic Representations of Uni- versal Laws, 387 Graphic Treatment of Arch, 481 Gravity, Acceleration of, 51, 79, 160 Gravity, Centripetal Force, 488 Gravity-Vertical, 488 Graphical Analysis, Elasticity, 489 Graphical Strengths, 497-498 Forces, 413-480 Guide Curve, 88 Gyroscope, 256 Harmonic Motion, 68, 81, 117 Isaac Newton's Laws, 68 Havardseus, Thales, Etc., 393-394 Height, Distance, 497 Holdkimen's Formulae for Columns,, 109 Horizontal Motion, 307 Hook's Law, 901, 305, 307 Hooke's Law, 901, 305, 307 Hanks, Structural Steel Beams, Horizontal Straints Girders by Graphical Method, 106-107 House-Power, 106-107 Lambeau's Law of Motion, 255-256 Lebanon's Law of Motion, 255-256 for Boat Speeds, 256-257 Impact, 68 Impact Law of Motion, 256-257 Insulated Beam, 497-498 Insulated Beam (Flexure), 497-498 Indicator Rods, 108 Inertia of Finito Rods, 59 Instantaneous Rotation, Axis of Rotation, Internal Stresses, General Prob- lem of Stress and Strain, Introd., 494 Lateral Axes, Lectropea, Kinetic Energy—Elasticity—Friction—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Force—Factorial, Lateral Security of Girders on Arch, Lever, Beam with Friction (see Lever), Lines and Curves, Live Loads, Load Lines, Loomotive on Arch, Loomotive on Girder, Loomotive on Rigid Rod, Malleable Cast Iron, Mass, Material Point, Mechanical Power and Heat No. Mechanics, Definition of, Mechanics Divisions of... Modulus of Elasticity... Modulus of Resilience... Modulus of Rigidity... Modulus of Tensacity... Mohr's Theorem... Moment of a Force... Moment Diagram... Moment of Inertia... Moment of Inertia Graphically Determined... Moment of Inertia of Box-sections... Moment of Inertia of Built-in Beam... Moment of Inertia of Built-in Column... Moment of Inertia of Built-in Figure... Bodies... Body of Truss... Moment of Torus... Momentum... Motion... Rotatory Motion... Rotatory Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotary Motion... Rotatory Moment (Flexure)... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Napierian Base,... Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Neutral Axis..., Normal Stress., Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stress Problem), Plane Stress Problem (see Plane Stresses and Shear Forces in Beams) Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Parallel Corde(s)... Pendulum Compound Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simple Circular Pendulum Simultaneous Reactions Pier Reactions 4 237 278 219 398 14 388 91 454 76 396 879 879 108 478 66 107 305 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 307 A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing the relationship between the neutral axis and the stress distribution in a beam. A diagram showing two parallel chords on opposite sides of a cross section. The chord on one side is labeled "parallel chord," while that on opposite side is labeled "parallel chord." The distance from each chord to its respective line is also shown. This represents how far apart two parallel chords can be placed within a cross section without affecting its strength. **INDEX TO MECHANICS OF SOLIDS.** (For index to "Hydraulics" see p. 304)
Piers of Arches 435, 438 Shafts, Non-Circular 389
Drilling 140 Shafts Diagram (Flexure) 965
Fall-Insulation 298 Shafting Distortion. 305, 308
(Square) Columna. 864 Shafting Distortion. 297
Pans, Vaulted um. Shafting Distortion in Plan um. 297
Polar Moment of Inertia 87, 336 Shafting Strains. 7, 825
Pole (in Graphs). 415, 417 Shafting Strains. 7, 825
Practical Noise. 135 Shearing Moment (Flexure). 305, 308, 304
Potential Power. 184 Shearing Strains. 7, 825
Power of Motor. 107, 187 Shearing Strains in the First & Second Order. 994
Power, Transmission of by Belting. 184 Slip (of Car, etc.) 161
Power, Transmission of by Shafts.
Spans of Shafts.
Equilibrium Polygons.
Specific Gravity.
Statica, Definition of.
Statica of a Material Point.
Statica of a Rigid Body.
Steam Engine Problems.
Steam Hammer.
Stone Strength of.
Stress Diagrams for Arch-Ribs.
Reduced Load-Contour Diagrams.
Relative and Absolute Velocity Diagrams.
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Resilience, Elasticity,
Index to Mechanics of Solids.
(For index to "Hydraulics" see p. xxxii)
Piers of Arches
(Square) Columna.
Pans, Vaulted
Polar Moment of Inertia
Pole (in Graphs).
Potential Power
Power of Motor
Belting
Power, Transmission of by Shafts.
Shafts.
Rays of Force Diagram.
Semiflexure Diagrams.
Reduced Load-Contour Diagrams.
Relative and Absolute Velocity Diagrams.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilence.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilience.
Elastic Resilence.
Elastic Resilience.
Elastic Resilence.
Elastic Resilience.
Elastic Resilence.
Elastic Resilience.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.
Elastic Resilence.                                                                           Shafts Diagram (Flexure).
Shafting Distortion.
Shafting Distortion in Plan um. Shafting Strains. Shearing Strains in the First & Second Order. Slip (of Car, etc.)
Steam Engine Problems.
Steam Hammer. Stone Strength of. Stress Diagrams for Arch-Ribs. Sudden Application of a Load. Summation of Products by Graphical Method. Table for Flexure. Table for Tension and Compression. Temperature Stresses. Tenacity Moduli. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. Thrust in Flexure. INDEX TO HYDRAULICS (For Index to " Mechanics of Solids" see p. xvii.)
Absolute Temperature Pearl Bernoulli's Theorem with Friction Pearl
Absolute Zero 806
Accumulator, Hydraulic 700 Billon, Experiments on Jets. 908
Adiabatic Expansion in Compressed Air Engine 885
Adiabatic Expansion in Gas from as Orifice 789 Borda's Formula. 725
Air Collector, Water 781, 798 Borda's Law. 615
Air Compressor, Transverse of of. 786-790 Bosanach Press. 526
Air compressor. 786-790 Bosyon's Petrolium engine. 796
Air pressure. 449 Boyant Effort of the Atmosphere. 608
Air pump, Sprengel's 636 Canal Locks, Time of Filling, 789, 740
Air pressure, Atmospheric as a Unit Pressure. 819 Continental Scale. 605
Amplitude of Backwater. 773 Centre of Pressure. 546
Anemometer, Birm. s. 887 Circular Cylinder. 415
Anemometer, Castello's. 887 Cody's Formula. 714
Angle of Head. 673 Circular Formula for Open Channels. 738
Angular Stiffness of Ship.



















































































<
Air Compressor, Transverse of of.
Air pressure.
Air pump, Sprengel's
Air pressure, Atmospheric as a Unit Pressure.
Amplitude of Backwater.
Anemometer, Birm. s.
Anemometer, Castello's.
Angle of Head.
Angular Stiffness of Ship.
Air Compressor, Transverse of of.
Air pressure.
Air pump, Sprengel's
Air pressure, Atmospheric as a Unit Pressure.
Amplitude of Backwater.
Anemometer, Birm. s.
Anemometer, Castello's.
Angle of Head.
Angular Stiffness of Ship.
Air Compressor, Transverse of of.
Air pressure.
Air pump, Sprengel's
Air pressure, Atmospheric as a Unit Pressure.
Amplitude of Backwater.
Anemometer, Birm. s.
Anemometer, Castello's.
Angle of Head.
Angular Stiffness of Ship. INDEX TO HYDRAULICS. xxiii (For index to "Mechanics of Solids" see p. xvi.) | Topic | Page | |------------------------------------------------------------------------|------| | Bernoulli's Theorem | 638 | | Contraction, - - - - - - - - - - - - - - - - - - - - - - - - - - | 678 | | Contracting Channels, Cooling in Sudden Expansion of Gas, Critical Temperature of a Vapor, Croton Aqueduct, Slope, Counting House, the River Ganges, Cope-steamers, Cups, Impulse of Jet on, Current, Transporting Power of, Curved Dams Thin, Holley Strength, Dams, High Mountains, Darcy, Experiments with Pliant's Tube, Darcy and Bernoulli with Open Channels, Decrease of Pressure along a Pipe, Depth of Water in a Pipe, Deviation of Spinning Ball from Vertical Plane, Diaphragm Valve, Pipe Loss of Head., Diaphragm Valve, Pipe Loss of Head., Diaphragm Valve, Divergent Tubes, Flow through Dividing Wall, Diving-bell., Double Fins for Ship's Drafts, Draught of Ships, Dubuca's Experiments, 707, 815., Duchemel's Formula for Wind-pressure., Duty of Pumping-engine., Dynamics of Gaseous Fluids, 715., Earthwork Dam., Earthwork Dam., Efficiency., Eiffel Tower., Eflux from Steam-boiler., Eflux into Condenser., Eflux under Water., Egg shape of Streamer for Engine., Elastic Fluids., Elbow, Length of Pipe., Ellipse, Book on Fire-streams., Empirical Law, Time of rise., Engine, Gas., Engine, Hot-air., Engine, Steam-, Enlargement, Sudden in Pipe., Equal Transmission of Fluid Pressure | 584 | | Equation of Continuity for Gases | 756 | | Equation of Continuity for Open Channels | 718 | | Erwin's Hot-air Engine | 756 | | Expanding Steam | 640 | | Fahrenheit Scale | 645 | | Fanning friction Coefficient | 605 | | Fanning friction Coefficient of Tubular Flow | 588 | | Fanning friction Coefficient of Fluid Friction | 709 | | Fire-engine Hose, Friction in Fire-streams by Ellis | 716 | | Flood gate | 538 | | Flow in Plane Layers | 632 | | Flow of Gas in Pipes | 768 | | Friction | 693, 707 | | Friction Coefficient | 828 | | Natural Gas | 534 | | Natural Gas and Equal Turbulence | 534 | | Natural Gas and Turbulence | 827 | | Francis' Formula for Overfall. Free Surface of Liquid at Rest. Fresh Water. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table. Heat. Hessian Table | 518 | | Frood's Experiments on Fluid Friction | 699 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 | | Frood's Experiments on Gradual Enlargement in Pipes | 725 xxiv INDEX TO HYDRAULIC. (For index to "Mechanics of Solids" see p. xvi.)
Gay-Lussac's Law 785 Jackson's Works on Hydraulics, 391 Page
Gradual Enlargement in Pipe. 785 Jet from Force-pump. 967
Gravities of Water. 785 Jet of Water. 900, 665, 938
Graphic Representation of 899 Jen, Experiment on Flow of 793
Change of State of Gas. 899 Water. 793
Loud of Water. 843 Kansas City Water-works: Si
Heat-engine Efficiency. 643 Kansas City Water-works: Si
Hessian's Law of Heat. 643 Kansas City Water-works: Si
Brevity of Fresh Water at 643 Kinetic Energy. 672, 719
Different Temperatures. 643 Kinetic Theory of Jet. 672, 719
Height Due to Velocity. 649 Kinetic Theory of Jets. 672, 719
Height of the Homogeneous
Page Number (xvii.)
Gay-Lussac's Law
Gradual Enlargement in Pipe.
Gravities of Water.
Graphic Representation of
Change of State of Gas.
Loud of Water.
Heat-engine Efficiency.
Hessian's Law of Heat.
Brevity of Fresh Water at
Different Temperatures.
Height Due to Velocity.
Height of the Homogeneous
Ammonia.
Herschel's Venturi Water-meter.
High Vacuum.
Hoop-tension.
Low Resistance to the Flow
time in...
Hot air.
Humphreys and Abbot's Survey
of the Manchester River.
Hutton's Formula for Wind
pressure.
Hydraulic Grade-line.
Hydraulic Head Depth.
Hydraulic Pressure.
Hydraulic Radius.
Hydraulic Power for Minimum
Friction.
Hydraulic Pressure.
Hydrodynamics.
Is Definition.
Hydrostatics.
Hydrostatic Pendulum.
Hydrostatics, Definition.
Joukowsky's Theorem.
Illuminating Gas.
Immersion of Rigid Bodies
Impurities in Water.
Impulse and Resistance of
Fluids in Motion.
Impulse of Jet on Vanea.
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
Page Number (xvii.)
') style_='display:none;'> id_='row_<'.concat('') data_index_row_id_='row_<'.concat('') data_index_col_id_='col_<'.concat('') data_index_page_id_=''.concat('') data_index_column_id_=''.concat('') data_index_table_id_=''.concat('') data_index_table_name_=''.concat('') data_index_table_position_=''.concat('') data_index_table_type_=''.concat('') data_index_table_sort_order_=''.concat('') data_index_table_sort_direction_=''.concat('') data_index_table_sort_field_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') data_index_table_sort_order_type_=''.concat('') INDEX TO HYDRAULICS (For index to "Mechanics of Solids" see p. xvi.) | page | page | |---|---| | Overfall, Emptying Reservoirs | 341 | | Overfall, Water, Actual Dia- change | 875 | | Overfall, Water, Actual Dia- change | 875 | | Overfall, Water, Actual Dia- change | 875 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 544 | | Parallelepipedal of a Steamer | 546 | | Parallelepipedal of a Steamer | 560 | | Pendulum, Hydrostatic. ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ........................................................................................................................................................................................................................?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?..?.. Pendulum, Hydrostatic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pendulum, Hydrostatic. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. Pendulum, Hydrostatic. .. .. .. .. .. .. .. .. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic. .. Pendulum, Hydrostatic... Pendulum, Hydrostatic. Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-ping Petroleum Gun-pinging. Pipe, Clos. Pipe, Four. Pipe, High-pressure. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Pipe Strength. Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in a Fluid Plate Moving in A... Plate Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to Flow Resistance to FlowResistance... Resistance of Water for Motion by Pressure Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motio... Resistance of Water for Motion by Pressure Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motio... Resistance of Water for Motion by Pressure Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motion by Pressure for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water for Motio... Resistance of Water fo... XXV xxvi INDEX TO HYDRAULICS. (For index to "Mechanics of Solids" see p. xiv.)
INDEX TO HYDRAULIC. For index to "Mechanics of Solids" see p. xviii.
Steam-engine, Examples. Vapour 315 607
Steam-gauge, Bourdon. Variable Diameter, Long Pipe, Page 794
Spring, Air Baggage. Long Pipe, Page 794
Stream-line. neta. 288
Bouyancy, Conducted Experiments. Velocity of Efflux as related to Density. 174
Sudden Diminution of Section. Velocity Measurements in Open 660
In a Pipe. Velocity Measurements in Open 730
Sudden Enlargement of Section. Vena Contracta. 639
In a Pipe. Vena Contracta. 639
Surface-Fluctuation. Venturi Tube, New Forma. 694
Survey, Hydraulics, of Mains. Venturi Tube, New Forma. 694
sight. Venturi Water-meter. 735
Tachometer.
Tachometer - Velocity Measurement - Absolute - Temperature Influence on Flow - Tension of Gas - Tension of Illuminating Gas - Thermometer - Thickness of Glass - Thickness of Pipe for Natumis - This Hollow Cylinder - This Piping - Throttle-valves, Loss of Head - Due to... - Time of Filling Vessels - Various Forms - Transom and Console - Air...
| Tachometer | Velocity Measurement | Absolute | Temperature Influence on Flow | Tension of Gas | Tension of Illuminating Gas | Thermometer | Thickness of Glass | Thickness of Pipe for Natumis | This Hollow Cylinder | This Piping | Throttle-valves, Loss of Head | Due to... | Time of Filling Vessels | Various Forms | Transom and Console | Air... | |-------------|----------------------|----------|----------------------------------|---------------|---------------------------------|------------|-------------------|-------------------------------|---------------------|----------------|---------------------------|---------|------------------------|--------------|------------------|-------------| | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | | (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is incomplete and some rows may be missing data.) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapour".) (Note that the table is complete but contains no information about "Vapoir". White background with no visible content.