Open PDF

You might also like

You are on page 1of 705
LZ Richard H. French OPEN-CHANNEL HYDRAULICS McGraw-Hill Book Company New York St.Louis San Francisco Auckland Bogota Hamburg Johannesburg London Madrid Mexico Montreal New Delhi Panama Paris $do Paulo Singapore Sydney Tokyo Toronto OPEN-CHANNEL HYDRAULICS INTERNATIONAL EDITION Copyright € 1986 Exclusive rights by McGraw-Hill Book Co — Singapore for manufacture and export. This book cannot be re-exported from the country to which it is consigned by McGraw-Hill 2nd Printing 1987 Copyright © 1985 by Richard H. French. All rights reserved. No part of this publication may be reproduced or distributed in any form or by any means. or stored in a data base or retrieval sytem, without the prior written permission of the publisher. The editors were Joan Zseleczky and Rita Margolies. The designer was Elliot Epstein The production supervisor was Sara L. Fliess. It was set in Century Schoolbook by University Graphics, Inc. Library of Congress Cataloging in Publication Data. French, Richard H Open-channel hydraulics. Includes index, 1. Channels (Hydraulic engineering) 1. Title. TCI75.F78 1985 627.1 85-207 ISBN 0-07-022134-0 When ordering this title use ISBN 0-07-Y66342-4 PAT lacie. sled SlelS | PRINTED AND BOUND BY B & JO ENTERPRISE PTE LTD, SINGAPORE Contents Preface, vii 1 Concepts of Fluid Flow 4 1.1 Introduction, 2 1.2. Definitions, 3 1.3. Governing Equations, 11 1.4 Theoretical Concepts, 24 1.5 Similarity and Physical Models, 37 2 Energy Principle 43 2.1 Definition of Specific Energy, 44 2.2 Subcritical, Critical, and Supercritical Flow, 47 2.3 Accessibility and Controls, 54 2.4 Application of the Energy Principle to Practice, 59 3 The Momentum Principle 75 3.1 Definition of Specific Momentum, 76 3.2. The Hydraulic Jump, 78 3.3 Hydraulic Jumps at Density Interfaces, 101 4 Development of Uniform Flow Concepts ANN 4.1 Establishment of Uniform Flow, 112 4.2 The Chezy and Manning Equations, 113 4.3 Resistance Coefficient Estimation, 115 5 Computation of Uniform Flow 163 5.1 Calculation of Normal Depth and Velocity, 164 5.2 Normal and Critical Slopes, 172 : 5.3 Channels of Composite Roughness, 176 5.4 Application of Uniform Flow Concepts to Practice, 182 6 Theory and Analysis of Gradually and Spatially Varied Flow 10 12 Desi: TA 7.2 7.3 = Basic Assumptions and the Equation of Gradually Varied Flow, 196 Characteristics and Classification of Gradually Varied Flow Profiles, 197 Computation of Gradually Varied Flow, 201 Spatially Varied Flow, 247 Application to Practice, 257 ign of Channels Introduction, 274 Design of Lined Channels, 279 Design of Stable, Unlined, Earthen Channels: A General Tractive Force Design Methodology, 288 Design of Channels Lined with Grass, 314 Flow Measurement Induction, 326 Devices and Procedures for Stream Gaging, 327 Structures for Flow Measurement: Weirs, 336 Structures for Flow Measurement: Flumes, 356 Structures for Flow Measurement: Culverts, 365 Rapidly Varied Flow in Nonprismatic Channels Turbi 10.1 10.2 10.3 10.4 10.5 Turbr 111 11.2 11.3 11.4 Introduction, 394 Bridge Piers, 394 Control of Hydraulic Jumps, 426 Drop Spillways, 441 Transition Structures, 444 ulent Diffusion and Dispersion in Steady Open-Channel Flow Introduction, 462 Governing Equations, 462 Vertical ani Transverse Turbulent Diffusion and Longitudinal Dispersion, 470 Numerical Dispersion, 499 Vertical, Turbulent Diffusion in a Continuously Stratified Environment, 501 ulent, Buoyant, Surface Jets and Associated Phenomena Introduction, 510 Basic Mechanics of Turbulent Jets, 511 Bouyant Surface Jets, 520 Upstream Cooling Water Wedges, 534 Gradually Varied, Unsteady Flow 12.1 12.2 12.3 124 12.5 12.6 Introduction, 550 Governing Equations and Basic Numerical Techniques, 554 Implicit Four-Point Difference Scheme—Channels of Arbitrary Shape, 560 Generalization of the Implicit Four-Point Technique, 572 Boundary and Initial Conditions, 574 Calibration and Verification, 576 195 273 325 393 464 509 549 CONTENTS vil 13. Rapidly Varied, Unsteady Flow ~ 583 14 13.1 13.2 13.3 13.4 13.5 Introduction, 584 Elementary Surges, 584 Dam Break, 594 Surges in Open Channels, 619 Pulsating Flow; Roll Waves, 621 Hydraulic Models 634 14.1 14.2 14.3 14.4 14.5 14.6 Introduction, 632 Fixed-Bed River or Channel Models, 638 Movable-Bed Models, 643 Model Materials and Construction, 656 Physical Model Calibration and Verification, 660 Special-Purpose Models, 661 Appendix I, 667 Appendix II, 673 Appendix III, 685 Index, 693 Preface From the twin viewpoints of quantity and quality, water-resources projects are of paramount importance to the maintenance and Progress of civilization as it is known today. The knowledge of open-channel hydraulics that is essential to water resources development and the preservation of acceptable water quality has, along with other areas of engineering knowledge, exploded in the last two decades. To some degree this explosion of information is due to the advent of the high-speed digital computer which has allowed the hydraulic engineer to address and solve problems which only 20 years ago would have been compu- tationally too large to contemplate. In addition, the concern of society with both the preservation and restoration of the aquatic environment has led the hydraulic engineer to consider open-channel flow processes which, previous to the environmental movement, would not have been considered important enough to require study. At this time there are only two well-known English- language books on~the subject of open-channel hydraulics—Open-Channel Hydraulics by V. T. Chow, published in 1959, and Open Channel Flow by F. M. Henderson, published in 1966. Neither of these books provides the modern viewpoint on this important subject. This book is designed primarily as a reference book for the Practicing engi- neer; however, a significant number of examples are used, and it is believed that the book can serve adequately as a text in either an undergraduate or graduate course in civil or agricultural engineering. As is the case with all books, this book is a statement of what the author believes is important, with the length tempered by a very reasonable page limit suggested by the publisher. Because of the background of the author, this book approaches open-channel hydraulics from the viewpoint of presenting basic principles and demonstrating the appli- cation of these principles. This book, unlike the ones that have preceded it, emphasizes numerical methods of solving problems; like previous hooks, the approach to the subject is primarily one-dimensional. X PREFACE The book is divided into 14 chapters. In Chapter 1 basic definitions and the equations which govern flow in open channels are introduced. In Chapters 2 and 3 the laws of conservation of energy and momentum are introduced, and applications of these basic laws to both rectangular and non- rectangular channels are discussed. In Chapters 4 and 5 uniform flow and its computation are discussed. In con- nection with the topic of uniform flow, flow resistance is also treated. In Chapter 6 gradually and spatially varied flow is considered. Solutions to these problems that involve tabular, graphical, and numerical procedures are considered. In Chapter 7 various methods of designing channels are presented. The types of channels considered include lined channels; stable, unlined, earthen chan- nels; and grass-lined channels. Consideration is given to both economics and seepage losses. In Chapter 8 methods of flow measurement are considered. Techniques included in this discussion are devices and procedures for stream gaging, weirs, flumes, and culverts. In Chapter 9 rapidly varied flow in nonprismatic channels is discussed. Dis- cussions of flow through bridge contractions; the control of hydraulic jumps with sharp- and broad-crested weirs, abrupt rises and drops, and stilling basins; drop-spillway structures; and the design of channel transitions are included. In Chapter 10 the transport processes known as turbulent diffusion and dis- persion are discussed. After the one- and two-dimensional governing equations have been developed, methods of estimating the vertical and transverse tur- bulent diffusion coefficients and the longitudinal dispersion coefficient are discussed. In Chapter 11 the topic of turbulent, buoyant surface jets in open-channel flov’ is discussed. In Chapters 12 and 13 gradually and rapidly varied unsteady flows are con- sidered. In these chapters modern numerical methods for solving the governing equations are discussed. In the case of the dam break problem, both numerical methods and simplified methods of computation are presented. In Chapter 14 the design, construction, and use of physical models in the study of open-channel hydraulics are discussed. Among the types of models considered are geometrically distorted and undistorted; fixed and movable bed; and ice. It should be noted and emphasized that this book is concerned only with the flow of water in channels where the water is not transporting significant quan- tities of air or sediment. In subject areas where pollutant transport is the pri- mary reason for describing the flow, specific pollutants are not discussed; rather, transport of a neutral tracer is considered. In the area of unsteady flow, tidal hydraulics are not considered since this is properly the venue of the coastal engineer. When the author began this book in the fall of 1981, he was, from the view- PREFACE xi point of 1984, not aware of the magnitude of the project he was undertaking. Because of his position as a member of the research faculty, the University of Nevada system could legitimately provide but verbal encouragement and moral support in this project. Thus, this project was completed in the evenings, on holidays, and on weekends. The author is especially indebted to Professor Vin- cent T. Ricca of the Ohio State University, who originally introduced the author to the subject of open-channel hydraulics, and the late Professor Hugo B. Fischer of the University of California, Berkeley, who continued the educa- tional process. The author would also acknowledge Dr. S. C. McCutcheon of the U.S. Geological Survey and Professor J. W. Bird of the University of Nevada, Reno, who spent numerous hours of their own time reading the manu- script. The author would also acknowledge his wife, Darlene, whose assistance through the seemingly endless and late hours of work provided invaluable encouragement that was desperately needed. Richard H. French OPEN-CHANNEL HYDRAULICS ONE Concepts of Fluid Flow 11 1.2 1.4 15 SYNOPSIS INTRODUCTION DEFINITIONS Types of Flow Changes of Depth with Respect to Time and Space Viscosity, Density, and Gravity Channel Types Section Elements Abstract Definitions GOVERNING EQUATIONS Conservation of Mass Conservation of Momentum Energy Equation Energy and Momentum Coetticients THEORETICAL CONCEPTS Sealing Equations Boundary Layers Velocity Distributions SIMILARITY AND PHYSICAL MODELS SYNOPSIS In Section 1.2, the types of flow encountered in open channels are clas- sified with respect to time. space, viscosity, density, and gravity. In addition, the types of channels commonly encountered and their geo- metric properties are defined. In Section 1.3, the equations which gov- ern flow in open channels, i.e., conservation of mass, momentum, and energy, are developed along with the equations for the energy and momentum correction coefficients. In Section 1.4, theoretical concepts such as the scaling of partial differential equations, boundary layers, and velocity distributions in both homogeneous and stratified flows are briefly discussed. In Section 1.5, the basic concepts of geometric and dynamic similarity are developed in relation to their application to the design and use of physical hydraulic models 1.1 INTRODUCTION By definition, an open channel is a conduit for flow which has a free surface, i.e., a boundary exposed to the atmosphere. The free surface is essentially an interface between two fluids of different density. In the case of the atmosphere, the density of air is much lower than the density for a liquid such as water. In addition the pressure is constant. In the case of the flowing fluid, the motion is usually caused by gravitational effects, and the pressure distribution within the fluid is generally hydrostatic. Open-channel flows are almost always turbulent and unaffected by surface tension; however, in many cases of practical impor- tance, such flows are density-stratified. The interest in the mechanics of open- channel flow stems from their importance to what we have come to term civi- lization. As defined above, open channels include flows occurring in channels ranging from rivulets flowing across a field to gutters along residential streets and continental highways to partially filled sewers carrying waste water to irri- gation channels carrying water halfway across a continent to vital rivers such as the Mississippi, Nile, Rhine, Yellow, Ganges, Amazon, and Mekong. Without exception, one of the primary requirements for the development, maintenance, and advancement of civilization is access to a plentiful and economic supply of water. In the material which follows, it is assumed that the reader is familiar with the basic principles of modern fluid mechanics and hydraulics, calculus, numer- ical analysis, and computer science. It is the purpose of this chapter to review briefly a number of basic definitions, principles, and laws with the focus being on their application to the study of the mechanics of open-channel flow. CONCEPTS OF FLUIDFLOW 3 1.2 DEFINITIONS Types of Flow Change of Depth with Respect to Time and Space As will be demonstrated in this section, it is possible to classify the type of flow occurring in an open channel on the basis of many different criteria. One of the primary criteria of classification is the variation of the depth of flow y in time ¢ and space x. If time is the criterion, then a flow can be classified as being either steady, which implies that the depth of flow does not change with time (éy/at = 0), or unsteady, which implies that the depth does change with time (dy/dt ¥ 0). The differentiation between steady and unsteady flows depends on the view- point of the observer and is a relative rather than an absolute classification. For example, consider a surge, i.e., a singular wave with a sharp front moving either up or down a channel. To a stationary observer on the bank of the channel, the flow is unsteady since he or she will note a change in the depth of flow with time. However, to an observer who moves with the wave front, the flow is steady since no variation of depth with time c-n be noted. If water is added or sub- tracted along the channel reach under consideration, which is the case with gutters and side channel spillways, then the flow may be steady, but it is non- uniform. Specifically, this type of flow is termed spatially varied or discontin- uous flow. If space is used as the classification criterion, then a flow can be classified as uniform if the depth of flow does not vary with distance (dy/dx = 0) or as nonuniform if the depth varies with distance (dy/ax # 0). Although concep- tually an unsteady uniform flow is possible, i.e., the depth of flow varies with time but remains constant with distance, from a practical viewpoint such a flow is nearly impossible. Therefore, the terminology uniform or nonuniform usu- ally implies that the flow is also steady. Nonuniform flow, also termed varied flow, is further classified as being either rapidly varied—the depth of flow changes rapidly over a relatively short distance such as is the case with a hydraulic jump—or gradually varied—the depth of flow changes rather slowly with distance such as is the case of a reservoir behind a dam. It should be noted that from a theoretical viewpoint the classifications of steady and uniform are very restrictive. For example, the terminology uniform flow implies that at every point in the flow field at an arbitrary instant in time, the velocity vectors have both the same magnitude and direction. Such a strict definition is much too restrictive for practical use; therefore, the definitions given above have been extended or relaxed to a point where they are useful. For example, in practice, time and space flow classifications are commonly done on the basis of gross flow characteristics. If the spatially averaged velocity of flow u = | | udA does not vary significantly with time, then the flow is 4 Depth Chence of depth trom time ee Te SEE = Sten * deptn | = Uniform Flow Unsteady Uniform Flow stuice Contraction below stuice gate QVF - Flood Wave Bore FIGURE 1.1 Various types of open-channel flow; GVF = gradually varied flow, RVF = rapidly varied flow. (Chow, 1959.) classified as steady. Similarly, if the average depth of flow is constant in space, then the flow is considered uniform. Figure 1.1 schematically provides examples of the above definitions applied to field situations. Viscosity, Density, and Gravity Recall from elementary fluid mechanics that, depending on the magnitude of the ratio of the inertial forces to the viscous forces, a flow may be classified as laminar, transitional, or turbulent. The basis for this classification is a dimensionless parameter known as the Reynolds num- CONCEPTS OF FLUID FLOW 8& ber, or r= UL (1.2.1) v where U = characteristic velocity of flow, often taken as the average velocity of flow L = characteristic length v = kinematic viscosity Then, a laminar flow is one in which the viscous forces are so large relative to the inertial forces that the flow is dominated by the viscous forces. In such a flow, the fluid particles move along definite, smooth paths in a coherent fashion. In a turbulent flow, the inertial forces are large relative to the viscous forces; hence, the inertial forces dominate the situation. In this type of flow, the fluid particles move in an incoherent or apparently random fashion. A transitional flow is one which can be classified as neither laminar nor turbulent. Jn open- channel flow, the characteristic length commonly used is the hydraulic radius which is the ratio of the flow area A to the wetted perimeter P. Then R =< 500 laminar flow 500 = R < 12,500 transitional flow 12,500 = R turbulent flow The state of flow based on the ratio of inertial to viscous forces is a critical consideration when resistance to flow is considered. Flows are classified as homogeneous or stratified on the basis of the variation of density within the flow. If in all spatial dimensions the density of flow is constant, then the flow is said to be homogeneous. If the density of the flow varies in any direction, then the flow is termed stratified. The absence of a density gradient in most natural open-channel flows demonstrates that either the velocity of flow is sufficient to completely mix the flow with respect to den- sity or that the phenomena which tend to induce density gradients are unim- portant. The importance of density stratification is that when stable density stratification exists, i.e., density increases with depth or lighter fluid overlies heavier fluid, the effectiveness of turbulence as a mixing mechanism is reduced. In two-dimensional flow of the type normally encountered in open channels, a commonly accepted measurement of the strength of the density stratification is the gradient Richardson number _ 8(4p/dy) i= plou/ay)? (1.2.2) where g = acceleration of gravity p = fluid density MANNE. AYO SCS ¥ = vertical coordinate du/dy: = gradient of velocity in vertical direction 9p/dy = gradient of density in vertical direction When u/dy’ is small relative to dp/dy, Ri is large, and the stratification is stable. When du/dv is large relative to dp/dy, Ri is small, and as Ri — 0, the flow sys- tem approaches a homogeneous or neutral condition. There are a number of other parameters which are used to measure the stability of a flow, and these will be discussed, as required, in subsequent chapters of this book. Depending on the magnitude of the ratio of inertial to gravity forces, a flow is classified as subcritical, critical, or supercritical. The parameter on which this classification is based is known as the Froude number F=—> (1.2.3) where U = a characteristic velocity of flow and L = a characteristic length. In an open channel, the characteristic length is taken to be the hydraulic depth, which by definition is the flow area A divided by the width of the free surface T, or . D=— (1.2.4) If F = 1, the flow is in a critical state with the inertial and gravitational forces in equilibrium. If F < 1, the flow is in a subcritical state, and the gravitational FIGURE 1.2 Propagation of an elementary wave. CONCEPTS OF FLUID FLOW 7 forces are dominant. If F > 1, the flow is in a supercritical state and the inertial forces are dominant. The denominator of the Froude number is the celerity of an elementary gravity wave in shallow water. In Fig. 1.2a an elementary water wave of height Ay has been created by the movement of an impermeable plate from the left to right at a velocity of Au. The wave has a celerity c, and the velocity of flow in front of the wave is zero. The situation defined by Fig. 1.2a is unsteady and cannot be analyzed by elementary techniques. However, as indicated in an ear- lier section in this chapter, some unsteady-flow situations can be transformed to steady-flow problems (Fig. 1.26). In this case, the transformation is accom- plished by adapting a system of coordinates which moves at a velocity c. This is equivalent to changing the viewpoint of the observer; i.e., in Fig. 1.2a the observer is stationary while in Fig. 1.2b the observer is moving at the velocity of the wave. Application of the steady, one-dimensional equation of continuity to the situation described in Fig. 1.2b yields cy = (y + ay)(c — Au) and simplifying, A c= yy (1.2.5) Application of the steady, one-dimensional momentum equation yields Byy? — bry + Ay)? = pey[(e — Au) — c] or 7 =8 , (1.2.6) Substitution of Eq. (1.2.6) in Eq. (1.2.5) yields 2 c or c= Vay (1.2.7) If it can be assumed that y ~ d (where d is the depth of flow), which is a valid assumption if the channel is wide, then it has been proved that the celerity of an elementary gravity wave is equal to the denominator of the Froude number. With this observation, the following interpretation can be applied to the sub- critical and supercritical states of flow: 1. When the flow is subcritical, F < 1, the velocity of flow is less than the celer- ity of an elementary gravity wave. Therefore, such a wave can propagate upstream against the flow, and upstream areas are in hydraulic communi- cation with the downstream areas. 2. When the flow is supercritical, F > 1, the velocity of flow is greater than the —Density intertace —-— ~- He ye KE KL 2 = =u te FIGURE 1.3 Notation for densimetric Froude number. celerity of an elementary gravity wave. Therefore, such a wave cannot prop- agate upstream against the flow, and the upstream areas of the channel are not in hydraulic communication with the downstream areas. Thus, the possibility of an elementary wave propagating upstream against the flow can be used as a criterion for differentiating between subcritical and super- critical flows. In the case of density stratified open-channel flows, it is often convenient to define an overall but inverted form of the gradient Richardson number. The internal or densimetric Froude number Fp is defined by p,- UY O” Velsolk Vel where, with reference to Fig. 1.3, Ap = p: — po, & = (0; — po)/p, and L = a characteristic length which is usually taken as the depth of flow in the lower layer y,. The interpretation of Fp is analogous to that of F; e.g., in an internally supercritical flow, a wave at the density interface cannot propagate upstream against the flow. (1.2.8) Channel Types Open channels can be classified as either natural or artificial. The terminology natural channel refers to all channels which have been developed by natural processes and have not been significantly improved by humans. Within this category are creeks, rivers large and small, and tidal estuaries. The category of artificial channels includes all channels which have been developed by human efforts. Within this category are nagivation channels, power and irrigation canals, gutters, and drainage ditches. Although the basic principles in this book are applicable to natural channels, a comprehensive understanding of flow in natural channels is an interdisciplinary effort requiring knowledge of several technical fields, i.e., open-channel hydraulics, hydrology, geomorphology, and sediment transport. Therefore, although flow in natural channels will be dis- cussed, an extensive treatment of the subject is outside the scope of this book. CONCEPTS OF FLUID FLOW 9 Since many of the properties of artificial channels are controlled by design, this type of channel is much more amenable to analysis. Within the broad category of artificial, open channels are the following subdivisions: 1, Prismatic: A prismatic channel has both a constant cross-sectional shape and bottom slope. Channels which do not meet this criterion are termed nonprismatic. 2. Canal: The term canal refers to a rather long channel of mild slope. These channels may be either unlined or lined with concrete, cement, grass, wood, bituminous materials, or an artificial membrane. 3. Flume: In practice, the term flume refers to a channel built above the ground surface to convey a flow across a depression. Flumes are usually constructed of wood, metal, masonry, or concrete. The term flume is also applied to lab- oratory channels constructed for basic and app'ied research. 4. Chute and Drop: A chute is a channel having a steep slope. A drop channel also has a steep slope but is much shorter than a chute. 5. Culvert: A culvert flowing only partially full is an open channel primarily used to convey a flow under highways, railroad embankments, or runways. Finally, it is noted that in this book the terminology channel section refers to the cross section of channel taken normal to the direction of flow. Section Elements In this part the properties of a channel section which are wholly determined by the geometric shape of the channel and the depth of flow are defined. 1. Depth of Flow y: This is the vertical distance from the lowest point of a channel section to the water surface. In most cases, this terminology is used interchangeably with the terminology depth of flow of section d, which is the depth of flow measured perpendicular to the channel bottom. The rela- tion between y and d is d cos 6 y= where @ is the slope angle of the channel bottom with a horizontal line. If @ is small yod Only in the case of steep channels is there a significant difference between y and d. It must be noted that in some places in this book y is also used to designate the vertical coordinate of a cartesian coordinate system. Although this dual definition of one variable can be confusing, it is unavoidable if a traditional notation system is to be used. 2. Stage: The stage of a flow is the elevation of the water surface relative to a datum. If the lowest point of a channel section is taken as the datum, then the stage and depth of flow are equal. 3. Top Width T: The top width of a channel is the width of the channel section at the water surface. 4. Flow Area A: The flow area is the cross-sectional area of the flow taken nor- mal to the direction of flow. 5. Wetted Perimeter P: The wetted perimeter is the length of the line which is the interface between the fluid and the channel boundary. 6. Hydraulic Radius R: The hydraulic radius is the ratio of the flow area to the wetted perimeter or A R=> 1.2. : P (1.2.9) 7. Hydraulic Depth D: The hydraulic depth is the ratio of the flow area to the top width or A D=—5 1.2. P (1.2.10) Table 1.1 summarizes the equations for the basic channel elements for the channel shapes normally encountered in practice. Irregular channels are often encountered in practice, and in such cases values of the top width, flow area, and the location of the centroid of the flow area must be interpolated from values of these variables, which are tabulated as a function of the depth of flow. Franz (1982) has developed a rational and con- sistent method for performing these interpolations. Define Ay = f, T. de (1.2.11) » y and YA, = J, (y — ¢)T, do = I A, do (1.2.12) where @ = dummy variable of integration T, = top width of channel at distance ¢ above origin (Fig. 1.4) A, = flow area corresponding to depth of flow y y = distance from origin to centroid of flow area Franz (1982) defined consistency of interpolation to mean: (1) both the tabu- lated and interpolated values are consistent with Eqs. (1.2.11) and (1.2.12), and CONCEPTS OF FLUID FLOW 44 (2) the variables vary monotonically between the tabulated values. Then, given tables of T and A as functions of y such that y; < ¥Y Sys: fori =0,1,..., n — l,and n + 1 = the number of depth values tabulated, T, = T;+ 2“ (7,,,— 7) (1.2.13) : Yi ~ Yi A, = A, + My — y(T, + T) (1.2.14) and = -YAy = FiAi + Hy — y(Ay + A) — Ky — yJ*AT, — T) (1.2.15) It must be noted that the accuracy of the values interpolated by Eqs. (1.2.13) to (1.2.15) hinges on the accuracy of the top width approximation because all other interpolated values are exact if the top width approximation is exact; con- versely, errors in the top width propagate to all other variables. Abstract Definitions A streamline is a line constructed such that at any instant it has the direction of the velocity vector at every point; ie., there can be no flow across a stream- line. A stream tube is a collection of streamlines. It is noted and emphasized that neither streamlines nor stream tubes have any physical meaning; rather, they are convenient abstractions. ’ 1.3. GOVERNING EQUATIONS In most open-channel flows of practical importance, the Reynolds number exceeds 12,500 and the flow regime is turbulent. Therefore, for the most part laminar flow regimes are not treated in this book. The apparently random nature of turbulence has led many investigators to assume that this phenome- non can be best described in terms of statistics. On the basis of this assumption, it is convenient to define the instantaneous velocity in terms of a time-averaged velocity and a fluctuating, random component. For a cartesian coordinate sys- tem, the instantaneous velocities in the x, y, and z directions are, respectively, u=ut+u’ veovt+u (1.3.1) and w=B+ uw’ Note: The average velocities used above may be determined by averaging either over time at a point in space or over a horizontal area at a point in time. In this book, the symbolism U7 indicates an average in time while @ indicates an average in space. It is essential that these definitions be recalled in later Channel section geometric elements | etted perimeter Channel type ! Area A P ~ T I i by | } - | | b + 2y | b+ 2yvTF2 \ zy? QyV1l +2 2 4 *Ty p+ oye +3 (d) . (8 — sin 8) 48 do *Satisfactory approximation when 0 < 4y/T < 1. TE ge 4 Pin ¥] For 4y/T > 1 P al l+lpi +i pt Ltlpl! Hyd Top width T by b by!s b+ 2y 7 ” (6 + zy)y b + 2zy (b + zy)y [(b + zy)y]"5 b+ w/t 22 b+ 2Qzy Vb + 2zy | T — 2 Qzy hy V2 as avi¢z | 2 7 | 2T*y 3A i ty 22 hy VET S'S aT? + ay 2y i ; | , | i | | i ' j | so 1, _sin8) aVxdo— 3) jL/O= sine) , | V2(0 — sin 4) we 4 6) 8) sin ko | | 32\V/sin 96 FIGURE 1.4 Definition of cross-sectional terms for interpolation. sections. From this point forward, the pertinent statistics will be defined in the x direction only with the tacit understanding that these definitions apply to the two remaining cartesian coordinate directions. The time-averaged velocity is defined as i a tr=7 di 3.) ur= a J, u t (1.3.2a) where T indicates a time scale which is much longer than the time scale of the turbulence. The spatially average velocity is given by a=4{{ dA * 1.3.26 wrasse (1.3.26) Then, since the turbulent velocity fluctuations are random in terms of a time average, 1 GG ? J u’ dt (1.3.3) Then the statistical parameters of interest are: 1. 2. 3. Root mean square (rms) value of the velocity fluctuations V2 17 rms(u’) = [3 J, (u’)? at| (1.3.4) T Jo Average kinetic energy (KE) of the turbulence per unit mass Average KE turbulence Mass Variable correlations measure the degree to which two variables are inter- dependent. In the case of the velocity fluctuations in the xy plane, the parameter | 1 fe u’v’ =F, u’v’ dt (1.3.6) measures the correlation which exists between u’ and v’. In a turbulent shear flow, u’v’ is finite; therefore, it is concluded that u’ and v’ are correlated. = 4)? + WW)? + Ww] (1.8.5) CONCEPTS OF FLUID FLOW 48 Conservation of Mass Independent of whether a flow is laminar or turbulent, every fluid flow must satisfy the equation of conservation of mass or, as it is commonly termed, the equation of continuity. Substituting the expressions for the instantaneous velocities defined in Eq. (1.3.1) into the standard equation of continuity (den- sity is assumed constant) (see, for example, Streeter and Wylie, 1975) yields a+ wu) | aw + v) 424 w) ox dy az ou au Ow du’ | du’ | dw’ a te a z =0 1.3.7) ax t ay * a2 t ox ay + a2 (1.3.7) or Taking the time or space average of each term in Eq. (1.3.7), ue, wm ox dy az aw ty aw" 0 ax | ay + Ge and Therefore, Eq. (1.3.7) yields two equations of continuity, one for the mean or time-averaged motion and one for the turbulent fluctuations oe ete ce (1.3.8) ox" ay" oz = ’ du’ | dv’ du’ d 5 +tp+ >= 1.3.9 . ar oy os : (2.3.9) In most applications of Practical importance in open-channel hydraulics, Eq. (1.3.8) is the only equation of continuity which is used. At this point and for the remainder of this section, the overbar and subscript which indicate a time average will be dropped. Other forms of the equation of continuity can be derived by specifying appropriate boundary conditions. For example, assume that the bottom bound. ary of the flow is given by the function F = y — yo(x,z) (1.3.10) contact with surface, the condition which must be satisfied at the bottom boundary is “ox Tay tus a0 (1.3.11) VO reir yee te eee Substitution of Eq. (1.3.10) into Eq. (1.3.11) yields 9¥0 4 4 249 _ . us +w Ge v(x,¥0,2) = 0 (1.3.12) In an analogous fashion, the free surface is defined by the function F=y— y*(x,2z,t) dy* ay* oy* ry +us ox +w Qe w(x,y*,z) = (1.3.13) where y* is the elevation of the free surface above a datum. Integrating Eq. (1.3.8) over the interval yo to y* and * ou y va yt [Mays |” May =0 | one » a poe ” du ” 8 yields I —dy+ oy + v(x,y*,z) — v(x,¥0,2) = 0 (1.3.14) yy Ox yy 92 Then a ay* ayo —_ = * _ =— re u dy = u(x,y*,2) = — ulx,¥o2) 3 du(x,y*,z) du(x,¥o2) + y* 7 2 Ox _ ox and it is noted that y ou du(x,y*,z) du(x,¥o2) Oe ay = ye DME 2) _ OU, Yor2) I. ax ax %0 ax Combining the above equations yields 8 y* a u dy = u(x,y* 2) — U(x,¥o,2 0 Bey, fee > dy x Jy, Y du 0 ° dyo oy* . a Po * 2) oy 01 ae dy az J», u dy + u(x,¥o,2) Ox u(x,y*,z) on (1.3.15) In an analogous fashion » ow ar” ayo _ I 32 dy 32 Jog w dy + w(x,¥,2z) a u(x,y’ ar (1.3.16) CONCEPTS OF FLUID FLOW 47 Combining Egg. (1.3.12) to (1.3.16) yields i ax J, Equation (1.3.17) is a form of the equation of continuity from which a number of forms used in the field of open-channel hydraulics can be derived. Some examples follow. arr ay* = == 1.3.17, udy +3 |, way + & 0 ( ) 1. Define » a= [ u dy » ‘0 y I w dy y 0 and a where q, and q, are the components of volume flux at the point (x,z) per unit time and unit widths of the vertical cross sections in the x and z directions, respectively, measured between the surface and bottom. With these defini- tions, Eq. (1.3.17) becomes . as 9% , ay ox ay ant 0 (1.3.18) 2. In many cases, it may be convenient to introduce values of u and w which are vertically averaged values, i.e., spatially averaged values. Define 1 fr and w= - w dy y* — Yo 4% Note: In this case u and w are average velocities in space rather than time. Rewriting Eq. (1.3.17) a wa [” | 2 [am [” | ay* sla} ua | +2 wdy|+2 =9 ox [= ee yy Jy, and substituting the definitions of u and w yields Ge Due i ee ox [Q* — yo)@] a a: a 7 (1.3.19) 3. In the case of a prismatic, rectangular channel Q = GA(t) = (y* — yb where b = channel width. Since the channel is prismatic, the flow can be considered one-dimensional, and therefore a . _ ay* = {["* - yo = 13: iF [(y you] + at 0 (1.3.20) aQ oy* or Pi +b a7 0 (1.3.21) In this rather lengthy discussion of the equation of continuity, an attempt has been made to derive the forms of this equation which will be utilized throughout this book and which are found in the modern literature. Conservation of Momentum A very useful approach to the derivation of an appropriate conservation of momentum equation for open-channel flow can be obtained from a consider- ation of the one-dimensional control volume in Fig. 1.5. The principle of con- servation of momentum states Rate of accumulation ) + of momentum within } | the control volume | ; net rate of c ‘ | | =5 momentum \ +4 sum of forces (1.3.22) | entering { | acting on control volume J control volume - The rate of momentum entering the control volume is the product of the mass flow rate and the velocity or Ax d Momentum entering = puy*(u) — a ae (puy*) FIGURE 1.5 One-dimensional control volume. CONCEPTS OF FLUID FLOW 49 Momentum leaving = puy*(u) + 2 2 (pu*y*) Then, the net rate at which momentum enters the control volume is = puy*(u) — = = 2 (put yt) = a = é ous 2 ou “y| = Ax ee (pu®y*) (1.3.23) The forces acting on the control volume shown in Fig. 1.5 are: (1) gravity, (2) hydrostatic pressure, and (3) friction. The body force due to gravity is the weight of the fluid within the control volume acting in the direction of the x axis or y = pgy* Ax sin 6 = pgy* Ax S, (1.3.24) where @ = angle the x axis makes with the bottom of the channel, and it is assumed that S, = sin@ The pressure force on any vertical section of unit width and water depth y* is oa | = |, pay= [" ceo — dy = Yoav? 0.8.25) The frictional force which is assumed to act on the bottom and sides of the channel is given by . F, = pgy* Ax S; (1.3.26) where S; = slope of the energy grade line or the friction slope. Combining Egg. (1.3.23) to (1.3.26) with the law of conservation of momen- tum, Eq. (1.3.22), yields ay"? a * a 2a * & —_ = _— - —=A: Ax 5 (ouy*) Ax 5 (ou'y*) + By*e(S — S)) 99% r Assuming that p is constant and dividing both sides of the above equation by pAx, dtuy*) ney , OY” y ea = *(S, — S, 1.3.27 at ou ) 2a, 7 & (S; /) ( ) which is known as the conservation form of the momentum equation. If this equation is soe du dy* du cee oy = — aap ae ant Og 2 (uy*) + ytu * a + gy* al = gyv(S,— S) (1.3.28) In an analogous fashion, the law of conservation of momentum can, when required, be written in a two-dimensional form or 8 yy 4 Zaye ¢ © uvysy + 89 (ye) = pytig. — 9 5p (uy d+ bx (uey*) + ay (uvy*) + 2 ’n (y*?) = gy*(S. — Sp) (1.3.29) and a a go 2 ae (vy*) + Bx (ue) +3 a 2 y*) +2 9 ay (y*)? = gy*(S, — Sp) (1.3.30) where S, and S, are the channel bottom slopes in the x and y directions, respec- tively, and S;, and S;, are the friction slopes in the x and y directions, respec- tively. These equations can also be expanded to yield forms similar to Eq. (1.3.28). In a prismatic, rectangular channel where the equation of continuity is given by Eq. (1.3.20), the corresponding conservation of momentum equation is du u2 oy* at 45 Ox At this point, it is convenient to note that the conservation of momentum in a turbulent flow is govemed by a set of equations known as the Reynolds equa- tions which can be derived from the Navier-Stokes equations by the substitu- tion of Eq. (1.3.1). The result in a general cartesian coordinate system is oe + es— = aS, — S) (1.3.31) au we, eae ae ta that Ba, + (ou, Fu i Hig t+ at ap] — 6 a db} | 8 | _ 80\ a (aD pig tat in te =- 2+ yh) +ul SS Ox oy az) oy \Ox a (1.3.32) . eo &) ou’ . av” . EN a0 — pl ” 2” Pax Ox oy dz Al? we, ww ow) _ 364 ne (ew ee 52 bekad _ a Mat Max + Pay + Mae) =~ a Pt tele eo FD aww awl’ aw” sa tsa! el a zs! oy’ dz \ Ox oy oz a where h = vertical distance p = density P = pressure Bw absolute or dynamic viscosity The above equations along with the law of conservation of mass govern all tur- bulent flows; but without an assumption to quantify the velocity fluctuations, a solution of this set of equations is not possible. CONCEPTS OF FLUIDFLOW 24 A popular method of quantifying the turbulent fluctuation terms in Eq. (1.3.32) is the Boussinesq assumption which defines the eddy viscosity such that the above equations become og t aoe tO + oe) 2 we amy tuto (+S og tag t OS tw =~ 2+ amy + n(F+ +S) (1.3.33) do _ ow wSC_ sD oe ela TUG TOS + BS) = — 5 + xh) (Pw : eo _ Fo) ax? " ay? * az? ] where » = eddy viscosity. It must be noted that the Boussinesq assumption treats 7 as if it were a fluid property similar to u; however, 7 is not a fluid prop- erty but is a parameter which is a function of the flow and density. In practice in turbulent flows, y is neglected since it is much smaller than nm Prandtl (see, for example, Schlichting, 1968), in an effort to relate the trans- port of momentum to the mean flow characteristics, introduced a characteristic length which is termed the mixing length. Prandtl claimed that +t = un~l’~e (1.3.34) ga dy where ¢ = mixing length. Then (1.3.35) where 7 = shear stress. A comparison of the Boussinesq and Prandtl theories of turbulence yields a relationship between the eddy viscosity and the mixing length or a and Tape, (1.3.36) p where 7/p = a kinematic turbulence factor similar to the kinematic viscosity. « is a direct measure of the transport or mixing capacity of a turbulent flow. In a 22 > OFEN-CHANNEL HYDRAULICS homogeneous flow, ¢ refers to all transport processes, e.g., momentum, heat, salinity, and sediment. In a density stratified flow, the following inequality is believed to be valid: e>&y > Ea (1.3.37) where cy = stratified momentum transport coefficient and «jy, = stratified mass transport coefficient. Energy Equation From elementary hydraulics and fluid mechanics, recall that the total energy of a parcel of fluid traveling at a constant speed on a streamline is equal to the sum of the elevation of the parcel above a datum, the pressure head, and the velocity head. The one-dimensional equation which quantifies this statement is known as the Bernoulli energy equation or = H = 24 + dacos6 + 2 (1.3.38) where subscript A = point on streamline in open-channel flow 2,4 = elevation of point A above an arbitrary datum d, = depth of flow section 6 = slope angle of channel u, = velocity at point A For small values of 6 Eq. (1.3.38) reduces to z H=z+yt 8 (1.3.39) where Z = spatially averaged velocity of flow and y = depth of flow Energy and Momentum Coeftticients In many open-channel problems it is both convenient and appropriate to use the continuity, momentum, and energy equations in a one-dimensional form. In such a case, the laws of conservation are: 1. Conservation of Mass (Continuity) Q=uA 2. Conservation of Momentum ZF = pQ(u, — %)) 3. Conservation of Energy We u H=z+yt oe CONCEPTS OF FLUID FLOW 23 Of course, no real flow is one dimensional; therefore, the true transfer of momentum through a cross section JI pu? dA A is not necessarily equal to the spatially averaged transfer pQu Thus, in situations where the velocity profile varies significantly in the vertical and/or transverse directions, it may be necessary to define a momentum cor- rection coefficient, or BpQu = If pu? dA 4 If pur dA Jf pu’ dA _ A — = a [7 pQu pu? A where 8 = momentum correction coefficient. In an analogous fashion, the kinetic energy correction coefficient is z 3 u u orgi= | [vzaa a : ff yu dA A Ivi | 1.3. and solving a ywA (1.3.41) and solving (1.3.40) where a = kinetic energy correction coefficient. The following properties of 3 and « are noted: 1, They are both equal to unity when the flow is uniform. In all other cases 8 and a must be greater than unity. 2.A comparison of Eqs. (1.3.40) and (1.3.41) demonstrates that for a given channel section and velocity distribution a is much more sensitive to the variation in velocity than 8. 3. In open-channel hydraulics 3 and a are generally used only when the channel consists of a main channel with subchannels and/or berms and floodplains (Fig. 1.6). In such cases, the large variation in velocity from section to section effectively masks all gradual variations in velocity, and it is appropriate to consider the velocity in each of the subsections as constant. In channels of compound section, the value of a may exceed 2. Sub-channet Main Channel FIGURE 1.6 Main channel with subchannels and floodplains. 1.4 THEORETICAL CONCEPTS Scaling Equations In general, the terminology of scaling of equations refers to a process by which a complex differential equation is examined in a rational manner to determine if it can be reduced to a simpler form given a specific application. The first step in the process is to nondimensionalize the equation. As an example, consider an incompressible, laminar, constant density and viscosity flow occurring in a gravity field. The Navier-Stokes equations along with the equation of conti- nuity govern the behavior of this flow. For simplicity, only the x component of the Navier-Stokes equations will be treated or du du du du : at + “ae tay + 3e dh ldap =-g- # dx pdx p\dx? — ay? Define the following dimensionless quantities: a 3 #u\ + ait oe) (1.4.1) a=Z o=5 e=5 i= a= (1.4.2) where L and U are a characteristic length and velocity, respectively. Substitut- ing these dimensionless variables in Eq. (1.4.1) yields Ut ou | U? aud | U* a6 | U? an Lat cat Tat Ta oh Uap | wU (3 eu aw Soe Lat * pLi\azi* at t ag) (143)

You might also like