River Engineering: John Fenton
River Engineering: John Fenton
John Fenton
Institute of Hydraulic and Water Resources Engineering Vienna University of Technology April 12, 2011
Table of Contents
Table of Contents 1 Introduction 1.1 The nature of what we will and will not do illuminated by some aphorisms and some people 1.2 The problem of ow in a river . . . . . . . . . . . . . . . . . . 2 River hydraulics 2.1 A note on terminology in the English language: 2.2 Summary . . . . . . . . . . 2.3 The one-dimensional equations of hydraulics 2.4 The behaviour of oods and waves in rivers . 2.5 Structures, controls, and boundary conditions 2.6 Computational hydraulics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 3 3 4 10 10 10 12 33 48 59 60 60 67
3 Measurement and analysis 3.1 Hydrometry and the hydraulics behind it . . . . . . 3.2 The analysis and use of stage and discharge measurements . 4 Sediment transport 4.1 General . . . . . . . 4.2 Initiation of motion . . . . 4.3 Bedforms and alluvial roughness 4.4 Transport formulae . . . . 4.5 Unsteady aspects . . . . 5 River morphology 5.1 Introduction . . . . . 5.2 Planform . . . . . 5.3 Longitudinal prole . . 5.4 Bends . . . . . . 5.5 Channel characteristics . . 5.6 Bifurcations and conuences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
78 . 78 . 78 . 78 . 78 . 78 79 . 79 . 79 . 79 . 79 . 79 . 79 80 . 80 . 80 . 80 . 80 . 80 . 80 81
6 River engineering 6.1 Introduction . . . . . . . . 6.2 Bed regulation . . . . . . . 6.3 Discharge control . . . . . . 6.4 Water level control . . . . . . 6.5 Water quality control . . . . . 6.6 River engineering for different purposes References
Chapter 1
Introduction
1.1 The nature of what we will and will not do illuminated by some aphorisms and some people
"There is nothing so practical as a good theory" stated in 1951 by Kurt Lewin (D-USA, 1890-1947): this is essentially the guiding principle behind these lectures. We want to solve practical problems, both in professional practice and research, and to do this it is a big help to have a theoretical understanding and a framework. We claim to aspire to deep connected simplicity rather than shallow disparate complexity. "The purpose of computing is insight, not numbers" the motto of a 1973 book on numerical methods for practical use by the mathematician Richard Hamming (USA, 1915-1998). That statement has excited the opinions of many people (search any three of the words in the Internet!). However, in contradiction to Hammings assertion, numbers are often important in engineering, whether for design, control, or other aspects of the practical world. A characteristic of many engineers, however, is that they are often blinded by the numbers, and do not seek the physical understanding that can be a valuable addition to the numbers. In this course we are not going to deal with many numbers. Instead we will deal with the methods by which numbers could be obtained in practice, and will try to obtain insight into those methods. Hence we might paraphrase simply: "The purpose of this course is insight into the behaviour of rivers; with that insight, numbers can be often be obtained more simply, cheaply, and reliably. "It is EXACT, Jane" a story told to the lecturer by a botanist colleague. The most important river in Australia is the Murray River, 2 375 km (Danube 2 850 km), maximum recorded ow 3 950 m3 s1 (Danube at Iron Gate Dam: 15 400 m3 s1 ). It has many tributaries, ow measurement in the system is approximate and intermittent, there is huge biological and uvial diversity and irregularity. My colleague, non-numerical by training, had just seen the demonstration by an hydraulic engineer of a computational model of the river. She asked: "Just how accurate is your model?". The engineer replied intensely: "It is EXACT, Jane". Nothing in these lectures will be exact. We are talking about the modelling of complex physical systems. A further example of the sort of thinking that we would like to avoid: in the area of palaeo-hydraulics, some Australian researchers made a survey to obtain the heights of oods at individual trees. This showed that the palaeoood reached a maximum height on the River Murray at a certain position of 18.01 m (sic), Having measured the cross-section of the river, they applied the Gauckler-Manning-Strickler Equation to determine the discharge of the prehistoric ood, stated to be 7 686 m3 s1 (sic) ... William of Ockham (England, c1288-c1348): "Occam"s razor is the principle that can be popularly stated as "when you have two competing theories that make similar predictions, the simpler one is the better." The term razor refers to the act of shaving away unnecessary assumptions to get to the simplest explanation, attributed to 14thcentury English logician and Franciscan friar, William of Ockham. The explanation of any phenomenon should make as few assumptions as possible, eliminating those that make no difference in the observable predictions of the explanatory hypothesis or theory. When competing hypotheses are equal in other respects, the principle recommends selection of the hypothesis that introduces the fewest assumptions and postulates the fewest entities while still suf ciently answering the question. That is, we should not over-simplify our approach. In general, model complexity involves a trade-off between simplicity and accuracy of the model. Occams Razor is particularly relevant to modelling. While added complexity usually improves the t of a model, it can make the model difcult to understand and work with. The principle has inspired numerous expressions including "parsimony of postulates", the "principle of simplicity", the "KISS principle" (Keep It Simple, Stupid). Other common restatements are: Leonardo da Vinci (I, 14521519, worlds most famous hydraulician, also an artist) his variant short-circuits the need for sophistication by equating it to simplicity "Simplicity is the ultimate sophistication". Wolfang A. Mozart (A, 17561791) "Gewaltig viel Noten, lieber Mozart", soll Kaiser Joseph II. ber die erste der groen Wiener Opern, die "Entfhrung", gesagt haben, und Mozart antwortete: "Gerade so viel, Eure Majestt, als ntig ist." (Emperor Joseph II said about the rst of the great Vienna operas, "Die Entfhrung aus dem Serail", "Far too many notes, dear Mozart", to which Mozart replied "Your Majesty, there are just as many notes as are 3
necessary"). The truthfulness of the story is questioned - Joseph was more sophisticated than that ... Albert Einstein (D-USA,1879-1955): "Make everything as simple as possible, but not simpler." Karl Popper (A-UK, 1902-1994) argued that a preference for simple theories need not appeal to practical or aesthetic considerations. Our preference for simplicity may be justied by his falsiability criterion: We prefer simpler theories to more complex ones "because their empirical content is greater; and because they are better testable". In other words, a simple theory applies to more cases than a more complex one, and is thus more easily falsiable. Popper coined the term critical rationalism to describe his philosophy. The term indicates his rejection of classical empiricism, and of the classical observationalist-inductivist account of science that had grown out of it. Logically, no number of positive outcomes at the level of experimental testing can conrm a scientic theory (Humes "Problem of Induction"), but a single counterexample is logically decisive: it shows the theory, from which the implication is derived, to be false. For example, consider the inference that "all swans we have seen are white, and therefore all swans are white," before the discovery of black swans in Australia. Poppers account of the logical asymmetry between verication and falsiability lies at the heart of his philosophy of science. It also inspired him to take falsiability as his criterion of demarcation between what is and is not genuinely scientic: a theory should be considered scientic if and only if it is falsiable. This led him to attack the claims of both psychoanalysis and contemporary Marxism to scientic status, on the basis that the theories enshrined by them are not falsiable. Falsiability, however, seems not so important a principle when it comes to engineering. Another objection is that it is not always possible to demonstrate falsehood denitively, especially if one is using statistical criteria to evaluate a null hypothesis. More generally it is not always clear, if evidence contradicts a hypothesis, that this is a sign of aws in the hypothesis rather than of aws in the evidence. However, this is a misunderstanding of what Poppers philosophy of science sets out to do. Rather than offering a set of instructions that merely need to be followed diligently to achieve science, Popper makes it clear in The Logic of Scienti c Discovery that his belief is that the resolution of conicts between hypotheses and observations can only be a matter of the collective judgment of scientists, in each individual case. Thomas Kuhn (USA, 1922-1996): In The Structure of Scienti c Revolutions argued that scientists work in a series of paradigms, and found little evidence of scientists actually following a falsicationist methodology. Poppers student Imre Lakatos (H-UK, 1922-1974) attempted to reconcile Kuhns work with falsicationism by arguing that science progresses by the falsication of research programs rather than the more specic universal statements of naive falsicationism. Kuhn argued that as science progresses, explanations tend to become more complex before a paradigm shift offers radical simplication. For example Newtons classical mechanics is an approximated model of the real world. Still, it is quite sufcient for most ordinary-life situations. Popper seems to have anticipated Kuhns observations. In his collection Conjectures and Refutations: The Growth of Scienti c Knowledge (Harper & Row, 1963), Popper writes, "Science must begin with myths, and with the criticism of myths; neither with the collection of observations, nor with the invention of experiments, but with the critical discussion of myths, and of magical techniques and practices. The scientic tradition is distinguished from the pre-scientic tradition in having two layers. Like the latter, it passes on its theories; but it also passes on a critical attitude towards them. The theories are passed on, not as dogmas, but rather with the challenge to discuss them and improve upon them." Another of Poppers students Paul Feyerabend (A-USA, 1924-1994) ultimately rejected any prescriptive methodology, and argued that the only universal method characterizing scientic progress was "anything goes!" The utility of deep simplicity compared with shallow complexity In view of the above, the lecturer prefers to approach problems with the former, deep simplicity, rather than the latter, shallow complexity. The following lecture notes will be a reection of that. Unfortunately "deep" sometimes looks complicated at rst, which can be off-putting. However, once a principle or a practice is established or understood, it seems simple in retrospect.
whose primary role is as a momentum diffusion or smoothing phenomenon, actually causes the ows to be unstable and to become turbulent. Most ows of interest to civil engineers are actually turbulent boundary layer ows.
1.2.1
Physical modelling
The problem of a turbulent ow with a free surface in the presence of irregular and moveable boundaries is so complicated that one effective way of modelling it is to build a physical scale model, which reproduces most of the real phenomena. This has been the traditional way of much hydraulics. There are problems, however, in that in a scale model, different physical properties of the uid and the ow have different relative importances, and the model is an approximate one. In the spirit of modelling, however, that is a legitimate approach and in hydraulics it has often provided answers where otherwise none could be obtained.
1.2.2
It is possible to apply computational uid mechanics (CFD) to solve such problems, which usually means the numerical solution in time and space of three-dimensional partial differential equations. Direct numerical simulation (DNS) is the most basic method where the Navier-Stokes equations are solved numerically in time, and the instability of the ow and the resulting turbulence is simulated. It captures all of the relevant scales of turbulent motion, so no model is needed for the smallest scales. This approach is extremely expensive, if not intractable, for complex problems on modern computing machines, hence the need for models to represent the smallest scales of uid motion. Reynolds-averaged NavierStokes equations (RANS) is the oldest approach to turbulence modeling. The velocities and pressure are written as a sum of slowly-varying plus rapidly-varying turbulent components. The equations are averaged mathematically over a time scale large compared with the turbulent uctuations, which gives equations for the slowly-varying terms, but which introduces new apparent stresses known as Reynolds stresses. It is necessary to introduce phenomenological models for such quantities. Such models include using an algebraic equation for the Reynolds stresses which include determining the turbulent viscosity, and depending on the level of sophistication of the model, solving transport equations for determining the turbulent kinetic energy and dissipation. Models include k (Spalding) and the Mixing Length Model (Prandtl). Large eddy simulation (LES) is a technique in which the smaller eddies are ltered and are modelled using a sub-grid scale model, while the larger energy carrying eddies are simulated. This method generally requires a more rened mesh than a RANS model, but a far coarser mesh than a DNS solution.
Vortex method is a grid-free technique for the simulation of turbulent ows. It uses vortices as the computational elements, mimicking the physical structures in turbulence.
1.2.3
Mathematical/theoretical modelling
For most hydraulics problems the irregularity of geometry, the large size, and the lack of data available, mean that there is an imbalance between the little that is known about a problem and the sophisticated methods that can be applied to simulate it. In this course we are going to take a different path, and use mathematical modelling. A better title might be "theoretical modelling". There is almost an inverse correlation between the sophistication of a model and the understanding that emerges: simple models provide insight, complicated models provide many results, but insight is harder to obtain. A mathematical model uses mathematical language to describe a system. Eykhoff in 1974 dened a mathematical model as "a representation of the essential aspects of an existing system (or a system to be constructed) which presents knowledge of that system in usable form". For our purposes, that is a quite useful denition. Often when engineers analyze a system to be controlled or optimized, they use a mathematical model. In analysis, engineers can build a descriptive model of the system as a hypothesis of how the system could work, or try to estimate how an unforeseeable event could affect the system. Similarly, in control of a system, engineers can try out different control approaches in simulations. Black-box and Clear-box modelling: Mathematical modelling problems are often classied into black box or white box (also called glass box or clear box) models, according to how much a priori information is available of the system. A black-box model is a system of which there is no a priori information available. A clear-box model is a system where all necessary information is available. Practically all systems are somewhere between the black-box and clear-box models, for which the term grey box can be used. Usually it is preferable to use as much a priori information as possible to make the model more accurate. Therefore the clear-box models are usually considered easier, because if one has used the information correctly, then the 5
model will behave correctly. In black-box models one tries to estimate both the functional form of relations between variables and the numerical parameters in those functions. Using a priori information we could end up, for example, with a set of functions that probably could describe the system adequately. If there is no a priori information we would try to use functions as general as possible to cover all different models. An often used approach for black-box models are neural networks which usually do not make assumptions about incoming data. The problem with using a large set of functions to describe a system is that estimating the parameters becomes increasingly difcult when the amount of parameters (and different types of functions) increases. This is usually (but not always) true of models involving differential equations. Any model which is not pure clear-box contains some parameters that can be used to t the model to the system it describes. If the modelling is done by a neural network, the optimization of parameters is called training. In more conventional modelling through explicitly given mathematical functions, parameters are determined by curve tting. As the purpose of modelling is to increase our understanding of the world, the validity of a model rests not only on its t to empirical observations, but also on its ability to extrapolate to situations or data beyond those originally described in the model. One can argue that a model is worthless unless it provides some insight which goes beyond what is already known from direct investigation of the phenomenon being studied.
1.2.4
Nature of problems
Input System Physical parameters or numerical coecients Physical equations or assumed equations
Nature of a physical system with input and output
Output
Figure 1.1:
Consider a physical system such as that shown in Figure 1.1, which is simplied here so as to show a single timedependent input, and also just one output quantity. In Clear-box modelling, such as the hydraulic model of a river, the model equations are known to some extent, and the physical parameters which are used by those equations are known, probably to a lesser extent. In a Black-box model, such as a neural network or a unit-hydrograph model of a catchment, the equations of the system are assumed generic response equations, linear or nonlinear, which are expressed in terms of numerical coefcients. Simulation is when both the system and the inputs are given, and it is required to calculate the responses, which might be, for example include, a water level hydrograph at a station, such as at Hainburg an der Donau. A common procedure in hydraulics is to assume values for physical parameters such as resistance of the stream (which might include, for example, telephoning your friend who worked on a similar river and asking her/him what they used for resistance ...). Rather more sophisticated is the numerical determination of physical (clear-box) or computational (black-box) parameters. This is the process of System Identi cation, where measured values of input and output are used to determine what the system characteristics are. This is also known as the solution of an Inverse problem. After the execution of this phase, then simulation can be performed with rather more condence.
Reference English Chanson, H. (1999), The Hydraulics of Open Channel Flow, Arnold, London. Chaudhry, M.H. (1993), Open-channel ow, Prentice-Hall. Chow, V.T. (1959), Open-channel Hydraulics, McGraw-Hill, New York. Francis, J.R.D. & Minton, P. (1984), Civil Engineering Hydraulics, fth edn, Arnold, London. French, R.H. (1985), Open-Channel Hydraulics, McGraw-Hill, New York. Henderson, F.M. (1966), Open Channel Flow, Macmillan, New York. Jain, S.C. (2001), Open-Channel Flow, Wiley. Julien, P.Y. (2002), River Mechanics, Cambridge.
Comments
Good introduction, also sediment aspects Very good readable but technical book Classic, now dated, not so readable Good elementary introduction Wide general treatment Classic, high level, readable High level, but terse and readable Ditto more applications to morphology
UBTUW JDF E222 JDF E222 JDF JDF E222 JDF UBTUW JDF JDF E222 UBTUW JDF JDF E222 JDF E222 JDF
Montes, S. (1998), Hydraulics of Open Channel Flow, ASCE, New York. Townson, J.M. (1991), Free-surface Hydraulics, Unwin Hyman, London. Vreugdenhil, C.B. (1989), Computational Hydraulics: An Introduction, Springer. Deutsch Forchheimer, Ph. (1930), Hydraulik, Teubner, Leipzig
Encyclopaedic & Denitive Simple, readable, mathematical Simple introduction to computational hydraulics
UBTUW
Naudascher, E., (1992), Hydraulik der Gerinne und Gerinnebauwerke, Springer Verlag, Wien, New York Preiler, G., Bollrich, G., (1985) Technische Hydromechanik, VEB Verlag fr Bauwesen, Berlin u Table 1.1: Introductory and general references
UBTUW UBTUW
Reference Boiten, W. (2000), Hydrometry, Balkema Bos, M.G. (1978), Discharge Measurement Structures, second edn, International Institute for Land Reclamation and Improvement, Wageningen. Bos, M.G., Replogle, J.A. & Clemmens, A.J. (1984), Flow Measuring Flumes for Open Channel Systems, Wiley. Fenton, J.D. & Keller, R.J. (2001), The calculation of streamow from measurements of stage, Technical Report 01/6, Cooperative Research Centre for Catchment Hydrology, Monash University. Novak, P., Moat, A.I.B., Nalluri, C. & Narayanan, R. (2001), Hydraulic Structures, third edn, Spon, London.
Comments A modern treatment of river measurement Good encyclopaedic treatment of structures Good encyclopaedic treatment of structures Two level treatment - practical aspects plus high level review of theory Standard readable presentation of structures JDF JDF E128
E222
Reference Cunge, J.A., Holly, F.M. & Verwey, A. (1980), Practical Aspects of Computational River Hydraulics, Pitman, London. Dooge, J.C.I. (1987), Historical development of concepts in open channel ow, in G. Garbrecht, ed., Hydraulics and Hydraulic Research: A Historical Review, Balkema, Rotterdam, pp.205 230. Flood Studies Report (1975), Flood Routing Studies, Vol. 3, Natural Environment Research Council, London. Lai, C. (1986), Numerical modeling of unsteady open-channel ow, in B. Yen, ed., Advances in Hydroscience, Vol. 14, Academic. Liggett, J.A. (1975), Basic equations of unsteady ow, in K. Mahmood & V. Yevjevich, eds, Unsteady Flow in Open Channels, Vol. 1, Water Resources Publications, Fort Collins, chapter 2. Liggett, J.A. & Cunge, J.A. (1975), Numerical methods of solution of the unsteady ow equations, in K. Mahmood & V. Yevjevich, eds, Unsteady Flow in Open Channels, Vol. 1, Water Resources Publications, Fort Collins, chapter 4. Miller, W.A. & Cunge, J.A. (1975), Simplied equations of unsteady ow, in K. Mahmood & V. Yevjevich, eds, Unsteady Flow in Open Channels, Vol. 1, Water Resources Publications, Fort Collins, chapter 5, pp. 183257. Price, R.K. (1985), Flood Routing, in P. Novak, ed., Developments in hydraulic engineering, Vol. 3, Elsevier Applied Science, chapter 4, pp. 129173. Skeels, C.P. & Samuels, P.G. (1989), Stability and accuracy analysis of numerical schemes modelling open channel ow, in C. Maksimovi & M. Radojkovi, eds, Computational Modelling c c and Experimental Methods in Hydraulics (HYDROCOMP 89), Elsevier.
Readable overview
Readable overview
JDF
Readable
JDF
The best overview of the advection-diusion approximation for ood routing Review
E222
JDF
Table 1.3: References on ood & wave propagation theoretical and computational
Reference The full equations for wave propagation and ood routing Cunge, J.A., Holly, F.M. & Verwey, A. ( 1980) Practical Aspects of Computational River Hydraulics, Pitman, London. Liggett, J.A. (1975) Basic equations of unsteady ow, Unsteady Flow in Open Channels, K.Mahmood & V.Yevjevich (eds), Vol.1, Water Resources Publications, Fort Collins, chapter 2. Liggett, J.A. & Cunge, J.A. (1975) Numerical methods of solution of the unsteady ow equations, Unsteady Flow in Open Channels, K.Mahmood & V.Yevjevich (eds), Vol.1, Water Resources Publications, Fort Collins, chapter 4. The advection-diusion approximation for ood routing Price, R.K. (1985) Flood Routing, Developments in Hydraulic Engineering, P.Novak (ed.), Vol.3, Elsevier Applied Science, chapter4, pp.129173. Dooge, J.C.I. (1986) Theory of ood routing, River Flow Modelling and Forecasting, D.A. Kraijenho & J.R. Moll (eds), Reidel, chapter3, pp.3965. Numerical methods fundamentals Liggett, J.A. & Cunge, J.A. (1975) Numerical methods of solution of the unsteady ow equations, Unsteady Flow in Open Channels, K.Mahmood & V.Yevjevich (eds), Vol.1, Water Resources Publications, Fort Collins, chapter 4. Noye, B.J. (1976) International Conference on the Numerical Simulation of Fluid Dynamic Systems, Monash University 1976, NorthHolland, Amsterdam; Noye, B.J. (1981) Numerical solutions to partial dierential equations, Proc. Conf. on Numerical Solutions of Partial Dierential Equations, Queens College, Melbourne University, 23-27 August, 1981, B.J. Noye (ed.), North-Holland, Amsterdam, pp.3137; Noye, B.J. (1984) Computational techniques for dierential equations, North-Holland, Amsterdam; Noye, J. & May, R.L. (1986) Computational Techniques and Applications: CTAC 85, North-Holland, Amsterdam. Smith, G.D. (1978) Numerical Solution of Partial Dierential Equations, Oxford Applied Mathematics and Computing Series, Second Edn, Clarendon, Oxford. Morton, K.W. & Baines, M. (1982) Numerical methods for uid dynamics, Academic; Morton, K. & Mayers, D. (1994) Numerical solution of partial dierential equations : an introduction, Cambridge; Morton, K.W. (1996) Numerical solution of convection-diusion problems, Chapman and Hall, London; Richtmyer, R.P. & Morton, K.W. (1967) Dierence Methods for Initial Value Problems, Second Edn, Interscience, New York.
Notes The best explanation of this eld A little disappointing, but the next best explanation
A more detailed introduction to nite dierence methods All are rather more comprehensive, describing some more general methods
Chapter 2
River hydraulics
2.2 Summary
Initially we consider the two equations for mass conservation, where almost no essential approximations are necessary: 1. Conservation of mass in the ow in the channel: one of the few exact equations in hydraulics (the hydrostatic pressure equation is another one). 2. Conservation of mass of sediment the Exner equation (Felix Maria von Exner-Ewarten , A, 1876-1930): also exact for uniform materials; the Wahl-Wiener who presents these lectures will defend this equation against attacks by foreign barbarians during the last 10 years. Then for momentum conservation, consider the hierarchy of river models, starting with the Navier-Stokes equations, where the models become simpler as we pass down the list: 1. The Navier-Stokes equations: the three-dimensional partial differential equations of uid mechanics; solving them is not feasible for most river problems. 2. Three-dimensional turbulent ow equations: where the turbulence is modelled. Solving them is only slightly more feasible for most problems. 3. Conservation of moment of momentum, allowing for vertical variation of pressure and velocity: introduced by Stefer in Canada, but it has never been exploited. One could develop a 2-D river model in which secondary currents were described, and which could simulate differential erosion on both sides of the river. 4. Boussinesq approximation: essentially a one-dimensional model, but which allows for variation of pressure across the ow due to curved streamlines. 5. Two-dimensional vertically-averaged river model: The real geometry of a meandering river can be included, 10
but no secondary currents are possible and erosion predictions are unreliable. 6. One-dimensional model using curvilinear co-ordinates: The long-wave equations for rivers that are curved in plan your lecturer ... 7. One-dimensional model for straight streams: the long-wave equations, shallow-water equations, "Saint-Venant" equations, the basis of most hydraulics. The solutions of these equations are much misunderstood, and some of the most elementary deductions and interpretations are wrong. 8. The low-inertia approximation: A rather simpler but still a good approximation 9. Advection-diffusion ood routing 10. Muskingum-Cunge routing 11. Kinematic wave routing 12. Black-box ood routing Figure 2.1 shows how all the theories relate to each other. The arrows generally show the direction of increasing approximation. The assumptions made in each case are shown as text without a box.
11
Real stream
Boussinesq approximations, non-hydrostatic, can describe transition between sub- and super-critical ow Assume pressure hydrostatic Two-dimensional equations, possibly including moment of momentum, to include secondary ows, not yet established Vertically averaged Two-dimensional equations, no secondary ows Assume stream curvature small One-dimensional long wave equations for curved streams, e.g. 2.3.1.3, eqn (2.4) Characteristic formulation, 2.4.3, reveals speed of front of disturbances Finite-dierence-based numerical methods, Preissmann Box Scheme Characteristic-based numerical models Assume small disturbances Linearised Telegraphers equation, 2.4.2, eqn (2.56), reveals nature of propagation of disturbances Assume ow and disturbances both slow Slow-change/slow-ow approximation, 2.4.5, eqn (2.79), simple equation & computational method Assume stream straight, no cross-stream variation One-dimensional long wave equations, eqns (2.22) and (2.23)
Both assumptions: small disturbances and slow ow Linearised, and slow-change/slow-ow, Advection-diusion equation, 2.4.6, eqn (2.82), reveals nature of most disturbances Assume diusion small Interchange of time & space dierentiation, Muskingum-Cunge routing Assume diusion zero Kinematic wave approximation
Figure 2.1:
12
In the past there has been little attention paid to dening co-ordinates or a control volume. Many derivations use as the fundamental space variable x some sort of distance along the channel bed, which is ambiguous for channels of arbitrary shape, and what is worse, this curved co-ordinate ideally requires special treatment and consideration of the curvature of the axis. It is easier instead to use cartesian co-ordinates, for which we use x the horizontal distance along the stream, y the horizontal transverse co-ordinate, and z the vertical, relative to some arbitrary origin.
2.3.1
A z Qb y x Ab Ab + Ab
A + A
Q + Q
Qb + Qb
Figure 2.2:
Consider Figure 2.2, showing an elemental slice of channel of length x with two stationary vertical faces across the ow. It includes two different control volumes. The surface shown by solid lines includes the channel cross section, but not the moveable bed, and is used for mass and momentum conservation of the channel ow. The surface shown by dotted lines contains the soil moving as bed load and extends down into the soil such that there is no motion at its far boundaries. Each is modelled separately, subject to a mass conservation equation, and each to a relationship that determines the ux. The uid in the channel might carry a suspended sediment load. Here, to good approximation, the concentration of suspended sediment will be considered constant so that the density of the uid-solid suspension is constant, the same as for any uid entering the control volume from rainfall, seepage, or tributaries, so that we can consider conservation of volume rather than of mass. Physically-oriented derivation This is a simpler and more intuitive derivation. Afterwards, a more mathematical derivation will be presented, as it provides something of a guide for when we consider momentum. On the upstream vertical face at any instant, there is a volume ux (rate of volume ow) Q, and on the downstream face Q + Q, so that Net volume ow rate of uid leaving across vertical faces = Q = Q 2 x + terms in (x) . x
If rainfall, seepage, or tributaries contribute an inow volume rate i per unit length of stream, the volume ow rate of this other uid entering the control volume is i x. If the sum of these two contributions is not zero, then volume of uid is changing inside the elemental slice, so that the water level will change. The rate of change with time t of uid volume inside the control volume can be expressed most simply in terms of rate of change of cross-sectional area as x A/t. Equating this to the net rate of uid entering the control volume, dividing by x and taking the limit as x 0 gives A Q + = i. (2.1) t x Remarkably for hydraulics, this is almost exact. The only approximation has been that the channel is straight. The second ow, the bed load, is composed of larger soil particles. The interstices between them are assumed to be occupied by the same uid as in the channel. In the absence of any detailed mechanism of transport or dilation, it 13
is assumed that the porosity of the bed load is constant. The composite bulk density of the bed load b is composed of the main solid constituent of the bed load, the water in the pores, and the suspended particles in the pore water, and it is assumed to be constant. The bed has a cross-sectional area Ab , the bulk volumetric ow rate is Qb , and there is an inow of mass rate mi per unit length, possibly due to deposition or erosion. Mass conservation is calculated following the same reasoning for this control volume as for that of the channel, giving Ab Qb mi . + = t x b (2.2)
The volume transport rate used here is the bulk ow rate; it is related to the volume ow rate of solid matter Qs used in transport formulae, by Qs = Qb (1 ), where is the porosity of the bed load. This is a slight generalisation of Exners law. Derivation using integral formulation with vectors There is insight to be gained by repeating the derivation using a general integral formulation for the conservation of any quantity. Here it will be considered initially for mass, later it will be used for momentum, when it is simpler to use this approach from the start. Consider the mass conservation equation for an arbitrarily moving and deforming control surface CS and control volume CV (White 2003, 3.2 & 3.3): Z Z d dV + ur dS = 0, n (2.3) dt CV CS {z } | {z } |
Total mass in CV
where t is time, dV is an element of volume, ur is the velocity vector of the uid relative to the local element of the control surface, which is possibly moving itself, n is a unit vector with direction normal to and directed outwards, and dS is an elemental area of the surface, as are shown on gure 2.3. The quantity urn = ur is the component n n
ur urn = ur n dS
Figure 2.3: The normal component of uid velocity relative to the control surface is urn = ur , hence the rate of volume n n transport is ur dS
of relative velocity normal to the surface at any point. It is this velocity that is responsible for the transport of any quantity across the surface, mass or momentum in our case. Now consider the control surface to be the elemental slice in gure 2.2, bounded by two xed vertical planes transverse to the channel, and bounded above by the free surface, which is free to move, and below by the soil surface, which is similarly free to move. An element of volume in the rst integral in equation (2.3) is dV = x dA, where dA is an element of cross-sectional area. The whole term becomes Z d A x dA = x , dt t
A
where it is necessary to use partial differentiation as A is a function of x as well. We have expressed the rate of change of mass in the control volume in terms of the rate of change of cross-sectional area. Considering the second integral in equation (2.3), on the free surface we have chosen the control surface to coincide with the free surface of the uid, so that particles on the free surface remain on the free surface, and the relative n normal velocity urn = ur 0 everywhere on that surface. Similarly particles on the interface between water and soil, or on any solid boundaries in the ow also dene the control surface, and there is also no contribution to mass or momentum transport. We include the usually poorly-known inow from rain, seepage or tributaries in an ad hoc manner, by just writing the contribution as x i, negative as it is an inow. 14
On the upstream face of the control surface, which is xed in space and through which uid moves, ur = u , n
where u is the x component of uid velocity, so that the contribution due to ow entering the control volume is Z u dA = Q,
A
Combining the contributions from the rate of change of mass in the control volume and the various contributions which cross the control surface, dividing by x and taking the limit x 0 we have the unsteady mass conservation equation A Q + = i, t x the same as equation (2.1) obtained by simpler but less-general means. Exactly the same procedure would be followed in the case of the bed-load, which would give equation (2.2). Streams curved in plan z Centre of curvature: n = r = 1/ n
where Q is the total volume ow rate or discharge in the channel. Similarly the downstream face contribution is + (Q + Q), where we use a Taylor expansion to write it as Q (Q + Q) = Q + x + . . . . x
nR
Figure 2.4:
Fenton & Nalder (1995) have considered streams curved in plan (i.e. almost all rivers!) and obtained the result: nm A Q + = i, (2.4) 1 r t s where nm is the transverse offset of the centre of the river surface from the curved streamwise reference axis s, which generally follows the path of the river and r is the radius of curvature of that axis. Usually nm is small compared with r, and the curvature term is a relatively small one, however for a highly meandering river there will be an effect on ood propagation velocities. Upstream Volume The mass conservation equation (2.1) suggests the introduction of a function V (x, t) which is the volume upstream of point x at time t, such that for the channel ow Z x V V i dx0 Q. (2.5) = A and = Q (x0 , t) + x t x0
The derivative of volume with respect to distance x gives the area, as shown, while the time rate of change of volume upstream is given by the rate at which the volume is increasing due to inow, minus the rate at which volume is passing the point and therefore no longer upstream. Substituting for A and Q into equation (2.1), shows that it is identically satised. The addition of Q (x0 , t) has been suggested by Fatemeh Soroush (Personal communication, 2011), and gives a denition which is an improvement on an earlier one used by the lecturer. In the case of the bed load, similarly introducing the upstream volume of the bed load Vb (x, t) such that Z x Vb Vb mi 0 (x ) dx0 Qb , (2.6) = Ab and = Qb (x0 , t) + x t b then the mass conservation equation (2.2) is identically satised. The use of these V and Vb will reduce the 15
total number of equations used from four to two, and enable the reduction of the combined problem to a single third-order differential equation. Use of stage instead of cross-sectional area We usually work in terms of water surface elevation stage , which is easily measurable and which is practically more important. We make a signicant assumption here, but one that is usually accurate: the water surface is horizontal across the stream. Now, if the surface changes by an amount in an increment of time t, then the area changes by an amount A = B , where B is the width of the stream surface. Taking the usual limit of small variations in calculus, we obtain A/t = B /t, and the mass conservation equation can be written B Q + = i. t x (2.7)
The discharge Q could be written as Q = U A, where U is the mean streamwise velocity over a section, and substituted into this. However, we believe the discharge to be more practical and fundamental than the velocity, and that will not be done here. When it comes to sediment transport, the velocity is important, but it can always be obtained from the discharge and the area.
2.3.2
The equation The conservation of momentum principle is now applied to the mixture of water and suspended solids in the main channel for a moving and deformable control volume (White 2003, 3.2 & 3.4): d dt Z u dV + Z u ur dS = P, n (2.8)
CV
CS
where u is the uid velocity, and P is the force exerted on the uid in the control volume by both body forces, which act on all uid particles, and surface forces which act only on the control surface. Note how similar the left side is to the left side of the mass conservation equation (2.3). It was a conservation equation for mass, integrating the mass per unit volume throughout the control volume and its ux across the control surface. Equation (2.8) is a conservation equation for momentum, obtained by integrating the contributions of uid momentum per unit volume, the quantity u. Contributions to force P 1. Body force: for the straight channel considered, the only body force acting is gravity; this force per unit mass is g, the gravitational acceleration vector. The total force on the uid in the control volume is the integral of this R with respect to mass, dm = dV , giving CV g dV = gV . We will consider only the x-component of the momentum equation, which have chosen to be horizontal, as g only has a component in the z direction, there will be no contribution from gravity to our equation! The manner in which gravity acts is to cause pressure gradients in the uid, giving rise to the following term, due to pressure variations around the control surface. 2. Surface forces: we separate these into pressure and shear forces, which act normally and tangentially to the surface respectively. a. Pressure the direction of the pressure force on the uid at the control surface is , where n is the n outward-directed normal; its local magnitude is p dS, where p is the pressure and dS an elemental area R of the control surface, as in gure 2.3. Hence, the total pressure force is the integral CS p dS. This n is difcult to evaluate for arbitrary control surfaces, as the pressure and the non-constant unit vector have to be integrated over all the faces. A simpler derivation is obtained if the term is evaluated using Gauss Divergence Theorem: Z Z p dS = p dV, n
CS CV
where p = (p/x, p/y, p/z), the vector gradient of pressure. This has turned a complicated surface integral into a volume integral of a simpler quantity. b. Shear forces: there is little that we can say that is exact about the shear forces. We will consider these when we make several practical assumptions.
16
Collecting terms Substituting the force contributions into the momentum equation (2.8) and taking just the x component: Z Z d u dV + u ur dS = n dt CV CS {z } {z } | |
Unsteady term Fluid inertia term
This expression, for an arbitrary control surface, has required no additional assumptions or approximations. Now, however, we consider the section of channel with the control surface as shown in gure 2.2. To obtain a useable expression it is necessary to make hydraulic approximations for all terms after the rst. Hydraulic approximations 1. Unsteady term The element of volume, as for the mass conservation equation, is dV = x dA, and the term contribution can be written Z d Q x u dA = x , (2.10) dt t where the integral A u dA has a simple and practical signicance it is just the discharge Q, so that the contribution of the term can be written simply as shown, but where again it has been necessary to use the partial differentiation symbol, as Q is a function of x as well. No additional approximation has been made in obtaining this term. It can be seen that the discharge Q plays a simple role in the momentum of the ow. 2. Fluid inertia term R The second term on the left of equation (2.8) is u ur dS, has its most important contributions from n
CS
(2.9)
the stationary vertical faces perpendicular to the main ow. If there is also uid entering or leaving from rainfall, tributaries, or seepage, there are contributions over the other faces, which we shall include below in an approximate manner. Considering these contributions separately: a. Stationary vertical faces: on the upstream face, ur = u, giving the contribution to the term of n R A u2 dA. The downstream face at x + x has a contribution of a similar nature, but positive, and where all quantities have increased over the distance x. The net contribution, the difference between the 2 two, after neglecting terms like (x) , can be written Z u2 dA. x x
A
In fact, the u term means that there are contributions from the turbulent uctuations. If we write u = u +u0 , where u is the mean value in time, and u0 is the turbulent uctuation, then substituting into u2 and taking the mean value in time, using the fact that u0 = 0 by denition, the term becomes Z u2 + u02 dA. x x
2 A
It can be shown that this is the only term in the momentum equation (2.9) where turbulence plays a role. To evaluate this integral it is necessary to have a detailed knowledge of the ow and turbulence distribution over the cross-section. As we have not solved the Navier-Stokes equations or the 3-D turbulent ow equations, that information is not available, and so here we make the rst major hydraulic assumption. The crude but simple approach is to assume that u2 and u02 are both proportional to the square of the mean velocity over the section (Q/A)2 , and to write the integral in terms of a Boussinesq coefcient : Z Q2 u2 + u02 dA . A
Traditionally the coefcient was used to make some allowance for the non-uniformity of velocity distribution. Fenton (2005) suggested that it should be modied to include also the effects of turbulence as shown. A typical value is = 1.05. The contribution to the equation is then simply but approximately. Q2 x . (2.11) x A 17
It will be shown below, that the whole term is of a relative magnitude that of the square of the Froude number, and its effects are small in many situations. It is useful to retain the , unlike many presentations that implicitly assume it to be 1.0, as it is a signal and reminder to us that we have introduced an approximation. b. Lateral momentum contributions: the remaining contribution to momentum ux is from inow such as tributary streams, or, although less likely, from ow seeping in or out of the ground, or from rainfall. In obtaining the mass conservation equation above, these were lumped together as an inow i per unit length, such that the mass rate of inow is i x, (i.e. an outow of i x) and if this inow has a mean streamwise velocity of ui before it mixes with the water in the channel, the contribution is x i i ui , (2.12)
where i is the Boussinesq momentum coefcient of the inow. The term is unlikely to be known accurately or to be important in most places. 3. Pressure gradient term the hydrostatic approximation R The contribution is CV p/x dV , the volume integral of the streamwise pressure gradient. The hydraulic approximation has a problem, because we have not attempted to calculate the detailed pressure distribution throughout the ow. However, in most places in most channel ows the length of disturbances is much greater than the depth, so that streamlines in the ow are only very gently curved, and the pressure in the uid is accurately given by the equivalent hydrostatic pressure, that due to a stationary column of water of the same depth. Hence we write for a point of elevation z, p = g Depth of water above point = g( z), where is the elevation of the free surface above that point. The quantity that we need is the horizontal pressure gradient p/x = g /x, and so the streamwise pressure gradient is entirely due to the slope of the free surface. The contribution is Z Z p dV x g dA x gA , (2.13) x x x
A A
where any variation with y has been ignored, as the surface elevation usually varies little across the channel, and so /x is constant over the section and has been able to be taken outside the integral, which is then simply evaluated. 4. Horizontal shear force term This term is probably the least well-known of all, and yet it is important in the equation. Other presentations assume that the slope of the channel is small, but that is unnecessary, and we shall not be forced to make it. The results will be applicable to steep spillways, inter alia. However, we are about to adopt a signicant approximate model of the process of resistance to motion. ASCE Task Force on Friction Factors in Open Channels (1963) recommended use of the Weisbach formulation for bed resistance, but that suggestion has been almost entirely ignored by the profession. The Gauckler-Manning-Strickler equation continues to be widely used, with the disadvantages noted by the Committee. Using the Weisbach formulation here provides additional insights into the nature of the equations and some convenient quantications of the effects of resistance. Consider the equation for the shear force on a pipe wall e.g. White (2003, 6.3): 2 (2.14) V , 8 where is the dimensionless Weisbach resistance factor, and V is the spatial and temporal mean velocity along the pipe. If the pipe is uniform and long, then is constant everywhere, both around the perimeter of the ow as well as along the pipe. The channel is a more complicated situation. Consider a vertical rectangular elemental prism shown in gure 2.5. The bed has a slope angle of in the x direction, which in many hydraulic problems is small. The transverse slope may be quite large however. The total shear force is s P , where s is locally downstream along the bed and P is an element of perimeter. This gives us Shear force on prismatic element = V 2 s P, (2.15) 8 in which the mean velocity tangential to the boundary V over the prism is assumed to be related to U , the mean x-component of velocity in the elemental prism, by V = U/ cos , as suggested in the gure. Now, the horizontal component of shear force, which is what we require, is obtained by multiplying the term (2.15) by cos = x/s, giving = Horizontal shear force on prismatic element = 18 2 V x P. 8
x Slope /x
U V
Element of channel showing dimensions and shear force from the bed
Summing around the perimeter, expressing it as an integral in perimeter distance, gives Z x Total horizontal shear force on control surface = V 2 dP. 8 P Now we make some gross approximations. We replace V , the mean velocity parallel to the bed, by U/ cos , as suggested in the gure, and if we write 1/ cos2 = sec2 = 1 + tan2 the integral becomes Z x U 2 1 + tan2 dP. Total horizontal shear force on control surface = 8 P As we usually do not know the variation of or the distribution of U or even in many cases the variation of the bed topography to give , we write it as the total perimeter P multiplied by a product of the mean values of the three terms in the integrand: 2 Z Q U 2 1 + tan2 dP 1 + S 2 P, A P where the quantities shown with tildes are mean values for the section: is the mean value of the dimensionless resistance factor around the perimeter, and the mean slope term S is the mean downstream slope evaluated around the wetted perimeter. In almost all river situations, where S is usually poorly known, it is so small that it plays no role here. In situations such as on steep spillways, where it is important, it is actually well-dened. Collecting terms gives x Q |Q| Total horizontal shear force on control surface = 2 1 + S 2 P, 8 A where the Q2 term have been replaced by Q |Q|, so that the direction of the stress is opposite to the ow, allowing also for tidal ow situations where the ow reverses: It is convenient for all subsequent work to introduce the symbol : = (2.16) 1 + S2 , 8 so that the result can be written Q |Q| (2.17) Total horizontal shear force on control surface = x 2 P. A Collection of terms and discussion Now all contributions from the hydraulic approximations to terms in equation (2.9) are collected, using equations (2.10), (2.11), (2.12), (2.13), and (2.17), and bringing all derivatives to the left and others to the right, all divided by x, gives the momentum equation: Q Q2 Q |Q| + i i ui (2.18) + + gA = P t x A x A2 19
At this stage the non-trivial assumptions in the derivation are stated, roughly in decreasing order of importance: 1. Resistance to ow is modelled by an empirical law, such that it is proportional to wetted perimeter and mean ow velocity squared. The law can incorporate nite effects of viscosity inasmuch as that affects the resistance coefcient, but no attempt is made to incorporate viscous or turbulent shear terms in the momentum balance. The Navier-Stokes equations are not being used. 2. All surface variation is sufciently long that the pressure throughout the ow is given by the hydrostatic pressure corresponding to the depth of water above each point. 3. Effects of curvature of the stream course are ignored. Its effects are of the order of the ratio of the width of the stream to its radius of curvature. 4. In the momentum ux term the effects of both non-uniformity of velocity over a section and turbulent uctuations are approximated by a momentum or Boussinesq coefcient. Nowhere else is any special treatment for turbulence necessary, so we have developed a model including turbulence. 5. Surface elevation across the stream is constant. 6. Suspended load concentration is a constant over the section at a particular time and place. In the resistance term, the only place necessary, allowance has been made for the stream being on a nite slope no limitation on slope has been required. Expressing A/x in terms of surface slope /x and mean bed slope S
B At x At x + x z A Z y Figure 2.6: Two channel cross-sections separated by x
In the momentum equation (2.18), if expanded, there would be a mixture of derivatives of area A/x and surface elevation /x. We choose to work with the more practical quantity of surface elevation, and in so doing, are forced to consider the bottom geometry in greater detail. The cross-section of a river in Figure 2.6 shows how ambiguous and possibly non-unique the concept of the bottom of the stream may be. In a distance x the surface elevation may change by an amount as shown, so that the contribution to the increase in cross-section area A is B , where is usually negative as the surface drops downstream. The change in the bed is Z, which in general varies across the section, with R contribution to A of B Z dy, the area between the solid and dotted lines on the gure corresponding to the bed at the two locations. The minus sign is because, if the bed drops away and Z is negative, as usual, the contribution to area increase is positive. Combining the two terms, Z A = B Z dy (2.19)
B
In practice the precise details of the bed are rarely known. For the latter contribution, the integral of the change in bed elevation across the stream, we introduce the symbol S for the mean downstream bed slope across the section such that Z 1 Z S= dy, (2.20) B x
B
where the negative sign has been introduced such that in the usual case when Z decreases with x, this mean downstream bed slope at a section is positive. In an important problem where bed details might be known, this can be evaluated. In the usual case where bed topography is poorly known, a reasonable local approximation or assumption is made. Using equations (2.19) and (2.20) we can write A = B + B S x, 20
where in a distance x the mean bed level across the channel then changes by S x under the water. In the rare case where the sides of the stream are vertical diverging or converging walls, an extra term would have to be included. Taking the usual calculus limit, we obtain A =B +S , (2.21) x x which might have been able to have been written down without the mathematical details.
2.3.3
In equation (2.18) weignore the last term, the inow momentum term i i ui as it is small and poorly known. Expanding the term Q2 /A /x, gives a term in /x which we also neglect, as we usually know little about how varies. There are now x-derivatives of both A and in the equation, which are not independent. We use equation (2.21) to eliminate rst one and then the other to give two alternative forms of the momentum equation governing ows and long waves in waterways. In both cases, we restate the corresponding mass conservation equation, using (2.1) and (2.7), to give the pairs of equations: Formulation 1 Long wave equations in terms of area A and discharge Q Eliminating /x gives the equations in terms of A and Q: A Q + = i, t x Q Q gA Q2 A Q |Q| Q . + 2 + 2 = gAS P t A x B A x A2 Formulation 2 Long wave equations in terms of stage and discharge Q Now eliminating A/x, but retaining A in all coefcients, as it can be calculated in terms of : 1 Q i + = , t B x B Q |Q| Q Q Q2 B Q2 B Q . + 2 + gA 2 = 2 S P t A x A x A A2 Nature of equations Each of the two formulations is in the form of a pair of equations that can be written as a vector evolution equation u u +C = r (u) , t x where u is the vector of unknowns, for example, [, Q], C is a 2 2 matrix with algebraic coefcients, and r is the vector of right side terms, due to inow, slope, and resistance. It can be shown that the system is hyperbolic, although this mathematical terminology seems not very useful for us. The implication of that is that solutions are of a wave-like nature. We will see that the behaviour of disturbances is more complicated than we might expect or is often stated. This arises because the right sides are functions of the dependent variables, that we have written here as r (u). In particular we will see that a common interpretation of the system in terms of characteristics, with the solution that of travelling waves with simple properties, is incorrect. The solution is actually more complicated: disturbances travel at speeds which depend on their length, and show diffusion as well. Comparison with previous presentations The questions arise, "how does this relate to other presentations of results?", and, "has the present approach actually done anything new?" It is strange, but there are almost no presentations in textbooks of the equations in this standard form. Most retain a term Q2 /A /x. As we have seen, the expansion of this is non-trivial. Textbooks also present results in terms of a depth-like quantity h (called "the depth"), which is presumably only useful for canal problems where the bottom is at, and the depth at the centre, can be dened. We will not deal with it, as boundary conditions are almost always expressed in terms of surface elevation . Presentations able to be recommended include those of Cunge, Holly & Verwey (1980) and Lai (1986). The above presentation has given a couple of additions, which might be useful in practice: 1. The use of the Boussinesq momentum coefcient , reminding us that the term is being modelled approximately only. Some (Cunge et al. 1980, Lai 1986) have included it, but also included equations without it. The 21 (2.23a) (2.23b) (2.22a) (2.22b)
lecturers claim that it is better to include it, rather than implicitly assuming it to be 1.0000 might be viewed cynically, when later in the lectures it will be suggested that terms in are of magnitude of the Froude number, and can usually be neglected anyway equivalent to making the = 0 "approximation"! 2. The use of mean downstream bed slope S, as dened in equation (2.20), in the right sides of the momentum equations (2.22b) and (2.23b): this is the only way in which variable bed topography has entered explicitly. Textbook presentations that do not expand the term Q2 /A /x do not confront this problem. In prac tice, however, S is often be calculated very approximately as, for example, even the water surface or even surrounding land elevation difference between two stations divided by the distance between them. 3. The resistance term: in equations (2.22b) and (2.23b) it appears as P Q |Q| /A2 ,which has a relatively simple interpretation as a dimensionless coefcient multiplied by the perimeter (over which the resistance acts), multiplied by the mean velocity in the stream, squared. This clearly follows from the approximate physical law assumed. There is no limitation on slope: these are long wave approximations, not shallow slope approximations. Alternative use of energy conservation The question also arises how would the use of energy conservation relate to this work? Consider the Energy conservation equation for the ow in the body of the channel (White 2003, 3.6): Z Z e dV + (p + e) ur dS = Losses, n (2.24) t
CV CS
where e is internal energy per unit mass of uid, which in hydraulics is the sum of potential and kinetic energies the e = gz + u2 + v 2 + w2 /2. Proceeding in a manner similar to the momentum equation above, we obtain 0 gB + t 2 t Q2 A + x Q3 Q2 |Q| , gQ + 1 2 = gHi i e P 2A A3 Z (2.25)
(u2 + v 2 + w2 ) un dA,
Hi is the head of the incoming ow, and e is a dimensionless energy loss coefcient. Equation (2.25) has two time derivative terms (which are clearly related to the rate of change of potential and kinetic energy at a section). It is not as simple as the momentum conservation equation, which is in terms of Q/t. To eliminate /t and A/t we use two different forms of the mass conservation equation, then, neglecting inow and 1 /x terms gives Q |Q| Q 0 + 31 Q Q Q2 B Q2 B . + + gA 1 2 = 1 2 S e P 0 t 2 A x A x A A2 This can be compared with the momentum equation (2.23b) for no inow: Q |Q| Q Q Q2 B Q2 B Q . + 2 + gA 2 = 2 S P t A x A x A A2 It can be seen that the structure of the two equations is the same, but they are slightly different in numerical detail. The leading term of the energy formulation is 0 Q/t; as 0 1.05, this gives a slightly different time scale from the momentum equation. As the coefcients 0 , 1 and are all just greater than 1, the numerical coefcients of the other terms are also just slightly different. In general, e is different from . It is the assertion in these lectures, however, that the processes of energy dissipation, as given by e , are more complicated than the boundary stresses that give rise to the coefcient . The momentum losses of the latter are a local phenomenon, to do with the local nature of the boundary and ow at the boundary only, whereas energy losses are more complicated and more distributed, in boundary layers, separation zones, vortices, and turbulent and viscous decay. Accordingly, the coefcient should be easier to quantify than e , and the momentum approach is to be preferred. It is also possible that the momentum response of the ow to boundary changes would be faster than energy processes, which consist of more complicated 3-D motions, and might take further downstream to adjust, as suggested in gure 2.7. The momentum model is simpler, and so, following Occams Razor, we are inclined to follow that path.
2.3.4
Figure 2.7:
For many centuries, if not millennia, the resistance to ow in open channels has claimed the attention of hydraulic engineers. And then: In this period of wide-spread scienti c development, a review of existing knowledge with regard to openchannel hydraulics might appear to be as lacking in timeliness as it is in glamour. Yet a glance at publications on the subject during the past decade or so will reveal many a signi cant anomaly. Momentum and energy analyses are at the same time over-simpli ed and confused with one another. Representation of parameters by symbols is mistaken for empirical formulation of functions. Flow formulas are sometimes said to involve the Froude number when they do not, and yet the Froude number is as frequently ignored when it is actually essential. Boundary texture and cross-sectional non-uniformity are often discussed without distinction. Roughness measures are mistaken for resistance coef cients and are permitted to re ect not only viscous effects but even changes in depth. In a word, principles of mechanics that have proved useful in other phases of uid motion are still not generally applied to problems of open-channel ow,which perforce remain subject to empirical solution all too often without even the guidance of physical reasoning. He then went on: However, it need only be recalled that the Chzy equation was unheard of two centuries ago the Manning type of formula, one century the Reynolds-number resistance diagram only a half century and the KrmnPrandtl relationships little more than a quarter century yet all are in regular use today. Were he alive today, he would despair, for the apparent convergence in development that he wrote about, has all seemed to zzle out. ASCE Task Force on Friction Factors in Open Channels (1963) wrote, warning about the use of the GaucklerManning-Strickler (G-M-S) formula: Many engineers have become accustomed to using Manning s n for evaluating frictional effects in open channels. At the present stage of knowledge, if applied with judgement, both n and are probably equally effective in the solution of practical problems. The design engineer who prefers to use n in her/his computations should continue to do so, but he/she should recognize the limitations on her/his method ... It is believed that experimental measurements of friction in open channels over a wide range of conditions are better correlated and understood by the use of . Furthermore, is commonly used by engineers in many other branches of engineering and probably provides the only basis for pooling all experience on frictional resistance in both open and closed conduits. It is recommended, therefore, that engineering teachers and research workers emphasize the use of the friction factor ... This has been almost completely ignored. What the Task Force was advocating was the use of the Weisbach approach, which we shall now consider. Dimensional analysis Consider the dimensional quantities governing the problem of the shear stress on the boundary of a stream which is straight: shear, where all quantities are taken to be section-averaged ones; the problem of variation around the boundary is too difcult for us at this stage: mean shear stress , discharge Q, cross-sectional area A, wetted
23
perimeter P , surface width B, mean roughness size k, gravitational acceleration g, density , and dynamic viscosity . From the 9 variables and 3 dimensions, from the Buckingham Theorem there are 6 dimensionless quantities. If we use as repeating variables , Q, and A, the remaining physical quantities whose importance we are to examine, are as shown in the rst column of table 2.2. The dimensionless results are shown in column 3. However in most cases the physical signicance of the terms is not clear. Multiplying by the factors in column 4 gives the dimensionless numbers in column 5, with more-easily-understood signicance. Dimensionless Factor with , Q, and A (Q/A)2 P/ A B/ A k/ A g A (Q/A)2 A Q 1 3
Physical quantity P
Signicance = , as introduced in equations (2.14) and (2.16) Aspect ratio of section: roughly width/depth
1 2
1/2
B/P
Top width / Bottom width Relative roughness: roughness size / mean depth normal to bed = 1/F 2 where F is the Froude number: eects of gravity = 1/R where R is the Reynolds number: eects of viscosity
5 6
1/3 2
Resistance coefcient : The dimensional analysis has shown how important is the quantity , that was introduced in equations (2.14) and (2.16). Traditionally, however, originating from pipe ows, the quantity considered is = 8, the Darcy/Weisbach coefcient. In pipe ows it is used as H = L V2 , D 2g (2.26)
where H is the head loss in a pipe of length L and diameter D. (There is a point of view that g plays no role in conned ow such as in a pipe and should not be used here instead, gH should be considered as a unity, the energy loss per unit mass of uid, such that the total rate of energy loss in the pipe is QgH). There is an interesting aside here, relating head loss and boundary shear stress. To do that, we actually use the integral momentum theorem applied to the control volume in a length of uniform pipe of length L inclined at an angle between stations 1 and 2, where here it is easier to consider a control volume with faces across the pipe rather than vertical as we did for the channel: D2 + (p1 p2 ) {z 4 } | D2 g L sin } | 4 {z = DL | {z } .
As L sin = z1 z2 we obtain
p1 p2 4 L z2 + = H = , z1 + g g gD
(2.27)
and so the quantity that we refer to as head loss has actually been obtained from momentum considerations. Comparing equations (2.26) and (2.27) we obtain equation (2.14) = 2 V . 8 24
It would probably have been more fundamental to have started by assuming = V 2 , the use of /8 came about because of using head and a circular pipe. Relative roughness k/(A/P ): It can be seen that the relative roughness has appeared in the form k/(A/P ), where we might think of the quantity A/P as the mean depth measured from the river bed, the important quantity in considering boundary resistance. Of course, many streams are wide, P B, such that A/P is equal to the mean depth A/B. Reynolds number R: Also in the last row of the table, the term which is the inverse of the Reynolds number, what appears as U D in many denitions of Reynolds number, velocity multiplied by transverse dimension, in this case is Q/P . Remaining quantities: the remaining quantities 2 , 3 and 5 are unimportant for pipe ows. For channel ows it is possible to use as a basis the extensive experimental results obtained in the early 20th century for pipes.
0.1
Completely turbulent
0.07
Relative roughness
0.001
0.0001
0.01
0.00001
103
104
107
108
Figure 2.8: The Moody diagram, showing graphically the solution of the Colebrook & White equation for as a function of R with a parameter the relative roughness = d/D.
The most important experimental results for industrial pipes were obtained by Colebrook and White (Colebrook & White 1937, Colebrook 1938). They obtained results for as a function of relative roughness = k/D, where D is pipe diameter (A/P = D/4 for circular pipes) and Reynolds number R = U D/. They presented an equation connecting , , and R, which was implicit in , the Colebrook-White equation. In the pre-computer age, that implicit relation was difcult to solve numerically. That has probably been the main reason that the approach has not been adopted. Moody (1944) plotted the results from that equation, as shown in Figure 2.8.There are, however, explicit solutions that are not as well-known as they should be. In channels, where the Reynolds number is usually so large that the ow is fully turbulent, Keulegan (1938)obtained for on the basis of a trapezoidal shape, 1 = log2 10 2.032
1 12.
k A/P
ln
1.3
1 k 12. A/P
(2.28)
where the rounded factor of 1/12. comes from a recommendation from ASCE Task Force on Friction Factors in Open Channels (1963). This simple approximation could have been the basis for much hydraulics, as it gives an explicit formula for the resistance factor in terms of the relative roughness k/(A/P ). Of course, determining the equivalent roughness k is a difcult problem, yet the knowledge of the manner of variation with roughness gives us an important tool. This is plotted in gure 2.9. It is notable just how little the resistance coefcient varies over a very wide range of roughnesses over a 100-fold roughness variation the resistance varies by a factor of only 3. A more general approximation has been obtained by Yen (1991, equation 30), including the results from a number of open channel experiments. Converting to natural logarithms here for simplicity, and rounding a coefcient 25
0.06
0.04 0.02
0 1/1000
Figure 2.9:
1/10
Fenton (2010) gave the formula here recommended for use in calculating the resistance coefcient: = ln2 1.33 0.9 . k 1 + 2. P 12. A/P Q (2.29)
Yen stated that such a formula was applicable for = k/ (A/P ) < 0.05 and R = Q/P > 30 000. If one does know the effective roughness size k, this provides the best way of calculating the resistance coefcient.
2.3.5
Steady uniform ow
The simplest model of a river is where the channel is prismatic, with a constant bed and surface slope S0 , and the ow is steady (/t = 0) and uniform (/x = 0). In equation (2.16) we introduced the symbol , where for typical streams, = /8, such that in the long wave equations (2.22b) or (2.23b), the resistance term appeared as P Q |Q| /A2 . In this case of steady uniform ow, /x = S = S0 , all terms except the gravity and resistance terms disappear from the momentum equation to give Q2 = gAS0 , (2.30) A2 giving the Weisbach-Chzy equation for the mean velocity U : r r Q gA A U= (2.31) = S0 = C S0 , A P P p p where we have introduced the Chzy coefcient C = g/ = 8g/. The most notable features are that the p velocity varies like A/P and like S0 . P There is an interesting deduction from equation (2.31). We calculate the square of the Froude number (widely used in open channel hydraulics) from it: U2 S0 B F2 = = , gA/B P
and as for wide rivers B P we have the conclusion that F 2 S0 /, and as S0 is constant, and to rst approximation is constant, so that Froude number is roughly constant for any given stream, independent of the ow rate. Gauckler-Manning-Strickler formula The resistance formulation most widely used in practice is the Gauckler-Manning-Strickler approach. To relate this to our formulation above, we consider its fundamental expression for a steady uniform ow: 2/3 p 2/3 p Q A 1 A U= S0 = kSt S0 , (2.32) = A n P P
in terms of the Manning coefcient n, which is dimensional, with units of L1/3 T, and the Strickler coefcient simply kSt = 1/n. The use of the symbol kSt for the Strickler coefcient is an unfortunate typographical coincidence with the mean roughness size k. Comparing this equation (2.32) with the Weisbach-Chzy equation (2.31), 26
the main difference is that here the velocity is modelled as varying like (A/P )2/3 , rather than (A/P )1/2 . For the two formulations (2.31) and (2.32) to agree, can be expressed as g = 2 = gn2 C P A 1/3 g = 2 kSt P A 1/3 . (2.33)
0.2
n g 0.1 k 1/6
0 1/1000
1/10
Figure 2.10: Theoretical variation of Mannings n with relative roughness, showing its relative insensitivity to depth as measured by A/P
Now using this relation we obtain a theoretical expression for the variation of Mannings n. Solving (2.33) and using = /8 gives s 1/6 A , (2.34) n= 8g P into which we substitute Keulegans expression for , (2.28), or Yens equation (2.29) with innite Reynolds number to give the theoretical expression for the variation of the apparent Manning n with relative roughness k/(A/P ): n g 0.40 (2.35) 1/6 . k 1/6 1 k k ln 12. A/P A/P This is plotted in gure 2.10, and it can be seen that over a wide range of relative roughnesses expected in nature, there is little variation of n g/k 1/6 ,leading to the widely-supposed justication for the G-M-S formula that the value of n is relatively invariant with depth. As an approximation, from the gure we might take the value of n g/k 1/6 to be constant at 0.125, leading to the approximate formula for n: k 1/6 n 0.125 . g (2.36)
Such an equation, with a numerical value of g implicitly substituted, is known as the Strickler equation, where k is taken to be the diameter of the bed material, such as d50 , the diameter of the bed material such that 50% of the 1/6 material by weight is smaller. If we substitute g = 9.8 ms2 we obtain n 0.040 d50 , where d50 is in metres, typical of the versions of the Strickler equation presented by French (1985, p160), but where some confusion follows from using d50 also in millimetres and feet. This enables us to propose a formula by substituting the Strickler-type formula (2.36) into the G-M-S expression (2.32) to give the formula for discharge in terms of known channel quantities: Q 1 U= = A 0.125 k 1/6 A P 2/3 p gS0 . (2.37)
In real rivers, however, n varies with depth it generally changes, as the nature of the roughness and the crosssection vary with depth. It is not a fundamental quantity it was not obtained from the boundary friction, although equation (2.34) shows how it can be so related. It has no theoretical justication it is an empirical approach which was convenient in a pre-computer age to to describe the problem approximately and conveniently, mimicking the behaviour of the Weisbach form. If we accept that in more general problems the resistance coefcient varies, then there is no real reason to use Mannings n rather than the more fundamental or . 27
The G-M-S formula is widely used and abused. The manner in which a value of n is adopted in practice is often most unsatisfactory. Although equation (2.36) does provides a basis for calculating the resistance coefcient from knowledge of the bed material, other more typical ad hoc approaches include: An Australian approach using the telephone (You did some work on River X years ago. What do you think n is on River Y (10km from X) for the reach between A and B?). Tables of values in books on open channels, given channel conditions, for example pp116-123 of Chow (1959), Barnes (1967) for US rivers, or Hicks & Mason (1991) for New Zealand Rivers. The photograph in gure 2.11 is of the Columbia River at Vernita in Washington, taken from Barnes. It has the lowest n = 0.024 of all the examples in the book. It is 500m wide and has boulders of diameter . . . well it doesnt say. Pity, because it is the roughness size to river depth which determines the resistance characteristics. Presenting a picture is not much help one has little idea of the underwater conditions and how variable they are.
Figure 2.11:
Conveyance and Friction Slope It is convenient to introduce the conveyance K (units of discharge L3 T1 ), so that the resistance term in the momentum equations appears as Q |Q| Q |Q| P = gA 2 . (2.38) A2 K From equation (2.33), the various denitions of K become K= r g A3 =C P r A3 A5/3 1 A5/3 = kSt 2/3 , = P n P 2/3 P (2.39)
such that in the form of equation (2.38) it is a convenient shorthand that includes the resistance coefcient of whatever law is being used, plus cross-sectional geometry terms. A simple and important result for steady uniform ow is that Q = K0 p S0 , (2.40)
where the subscript 0 denotes the conveyance corresponding to the normal depth h0 of the uniform ow. Many textbook presentations write the friction term in terms of a dimensionless quantity Sf = Q |Q| /K 2 , called the "friction slope", possibly better known as "resistance slope", so that the resistance term in the momentum equations appears as gASf . It is also known as the "slope of the energy grade line", or the "head gradient", which gives an uninformative and misleading picture, for in our momentum-based approach it is neither of those things. Computations of steady uniform ow As we know, steady ow does not change with time; uniform ow is where the depth does not change along the waterway. For this to occur the channel properties also must not change along the stream, such that the channel is prismatic, and this occurs only in constructed canals. However in rivers if we need to calculate a ow or depth, it is common to use a cross-section which is representative of the reach being considered, and to assume it constant for the approximate application of theory. This is the simplest problem we consider! The Weisbach and Chzy equations and the Gauckler-Manning-Strickler forms (2.31) and (2.32) are formulae for 28
in which both A and P are functions of the ow depth. Each equation show how ow increases with cross-sectional area and slope and decreases with wetted perimeter. If we know or can assume the resistance parameter, this is straightforward. However if we have to calculate it from the Yen formula (2.29), which is in terms of Q, it has to be done iteratively see example 2.1 below. Trapezoidal sections B h P W
Figure 2.12: Trapezoidal section showing important dimensions
the discharge Q in terms of resistance coefcient, slope S0 , area A, and wetted perimeter P : r r 1/2 p 8g A 8g A S0 = A S0 Weisbach-Chzy : Q=A P P 2/3 p A A S0 G-M-S : Q= n P
(2.41a) (2.41b)
Most canals are excavated to a trapezoidal section, and this is often used as a convenient approximation to river cross-sections too. In many of the problems in this course we will consider the case of trapezoidal sections. We will introduce the terms dened in Figure 2.12: the bottom width is W , the depth is h, the top width is B, and the side slope, unusually dened to be the ratio of H:V dimensions is . From these the following important section properties are easily obtained: Top width : Area : Wetted perimeter : (Exercise: Obtain these relations). The hydraulically-wide channel Many rivers and canals are sufciently wide that we can neglect the effects of the sides and for purposes of simple calculations we say A Bh and P B, so that we model the section as a rectangle that is much wider than it is deep, and it is called hydraulically-wide. We introduce the discharge per unit width q = Q/B, which is equal to the integral of velocity over the depth. In this case equations (2.41a) and (2.41b) become r 8g 3/2 p Weisbach-Chzy : q= S0 and (2.43a) h p 1 G-M-S : q = h5/3 S0 , (2.43b) n and we see that q h1.5 or h1.67 . An immediate deduction is, for example, that if a ow increases by a factor of 10, the the depth increases by a factor of approximately 102/3 4.6, while a 100-fold increase in ow means an increased depth with a factor of 1002/3 22. Computation of discharge This is done by evaluating any of the equations. If the numerical value of the resistance coefcient is to be calculated from the Yen formula (2.29) it has to be done iteratively. Example 2.1 Calculate the ow in a river of slope S0 = 103 , approximated by a trapezoidal section with bottom width
10 m, side slopes 2 : 1, and depth 1 m, if the mean roughness size is 1 cm, using (a) the Yen formula (2.29) and the Weisbach-Chzy formula (2.41a) we use the viscosity of water at 20 C = 1 106 m2 s1 . (b) the Strickler formula (2.36) and the G-M-S formula (2.41b) combined to give equation (2.37) above With h = 1 m, equation (2.42b) gives A = 1 (10 + 2 1) = 12.0 m2 , equation (2.42c) gives P = 10 + 2 1 + 22 1 = 14. 47 m.
B = W + 2h A = h (W + h) p P = W + 2 1 + 2 h.
29
(a) Weisbach Firstly we calculate the , neglecting the viscosity term = ln2 Calculate Q Q=A 8g A S0 = 12.0 P 8 9.8 12 0.001 = 18.3 m3 s1 0.0279 14.47 1.33 ln2
1 0.01 12. 12/14.47 14.47 + 2. 110 18.3
6
1.33
k 1 12. A/P
+ 2. P Q
0.9
1.33 ln
2 0.01 1 12. 12/14.47
= 0.0279
Repeat (i) and (ii), including that result in the viscosity term =
0.9
= 0.0280,
and we see that the viscosity correction is not worth incorporating. (b) G-M-S/Strickler/Equation (2.37) Q= A 0.125 k1/6 A P
2/3
gS0 =
12 0.125 0.011/6
12 14.47
2/3
Computation of normal depth If the discharge, slope, and the appropriate roughness coefcient are known, equations (2.41a)-(2.41b) is a transcendental equation for the normal depth h, which can be solved by the methods for solving transcendental equations. A simple method is to use direct iteration, whereby we arrange the equations in the form h = f (h), and solve by repeated evaluation. For this to converge, f (h) must be a relatively slowly-varying function. Here we develop a procedure which seems to work well: In the case of hydraulically-wide channels, we neglect the contributions of the side slopes so that neither the wetted perimeter P nor the breadth B vary much with depth h. Hence the quantity A(h)/h also does not vary strongly with h. Hence we can rewrite the G-M-S expression: 1 A5/3 (h) p S0 (A(h)/h)5/3 Q= h5/3 , S0 = 2/3 (h) nP n P 2/3 (h)
where most of the variation with h is contained in the last term h5/3 , and by solving for that term we can re-write the equation in a form suitable for direct iteration 3/5 Qn P 2/5 (h) h= , A(h)/h S0 where the rst term on the right is a constant for any particular problem, and the second term is expected to be a relatively slowly-varying function of depth, so that the whole right side varies slowly with depth a primary requirement that the direct iteration scheme be convergent and indeed be quickly convergent. Here, in addition, we include the Strickler formula (2.36), which is equation (2.37) re-written 3/5 P 2/5 (h) Q k 1/6 . h = 0.125 A(h)/h S0 g
(2.44)
Experience with typical trapezoidal sections shows that this works well and is quickly convergent. However, it also works well for ow in circular sections such as sewers, where over a wide range of depths the mean width does not vary much with depth either. For small ows and depths in sewers this is not so, and a more complicated method such as the secant method might have to be used. Example 2.2 Calculate the normal depth in the previous example 2.1, for the calculated discharge of 18.1 m3 s1 using the
G-M-S formulation, with equation (2.44). We pretend we do not know the answer of 1 m and for an initial estimate we choose 1.5 m. We have A = h (10 + 2 h), P = 10 + 4.472 h, giving the scheme h = = = Q k1/6 0.125 S0 g
3/5
30
and starting with h = 1.5 we have the sequence of approximations: h = 0.979 m, h = 1.002 m, h = 1.001 m, and the method has converged to this result, quite satisfactorily in its simplicity and speed (it did not converge to the exact result because the specied discharge was rounded).
2.3.6
Steady gradually-varied ow
A common problem in river engineering is, for example, how far upstream water levels might be changed, and hence ooding possibly enhanced, due to downstream works such as the installation of a bridge or other obstacles. Far away from the obstacle or control, the ow may be uniform, but generally it is variable. The transition between conditions at the control or point of known level, and where there is uniform ow is described by the GraduallyVaried Flow Equation, which is an ordinary differential equation for the water surface height. The solution will approach uniform ow if the channel is prismatic, but in general we can treat non-prismatic waterways also. The steady ow approximation is often used as a rst approximation, even when the ow is unsteady, such as in oods. The differential equation Consider the mass conservation equation for steady ow, when /t 0, and equation (2.1) becomes dQ = i, dx with the solution obtained by integration Q(x) = Q0 + Z
x
i dx,
x0
where at an upstream station x0 the discharge is Q0 , the extra discharge just being given by the integral of the inow i. The momentum equation (2.23b) for Q/t = 0 and Q/x = i, and assuming Q positive, becomes Q2 B d Q Q2 Q2 B gA 2 = 2 S P 2 2 i, A dx A A A
(2.45)
which is a rst-order ordinary differential equation for (x), provided we have evaluated Q(x), and that we know how the geometric quantities A and B depend on surface elevation at each point. This is the Gradually-Varied Flow Equation. The equation may be solved numerically using any of a number of methods available for solving ordinary differential equations which are described in books on numerical methods. These methods are usually accurate and can be found in many standard software packages. It is surprising that books on open channels do not recognise that the problem of numerical solution of the gradually-varied ow equation is actually a standard numerical problem, although practical details may make it more complicated. Instead, such texts use methods such as the Direct step method and the Standard step method, which can become complicated. There are several software packages such as HEC-RAS which use such methods, but solution of the gradually-varied ow equation is not a difcult problem to solve for specic problems in practice if one recognises that it is merely the solution of a differential equation. In sub-critical (relatively slow) ow, the effects of any control can propagate back up the channel, and so it is that the numerical solution of the gradually-varied ow equation also proceeds in that direction. On the other hand, in super-critical ow, all disturbances are swept downstream, so that the effects of a control cannot be felt upstream, and numerical solution also proceeds downstream from the control. No inow: If there is no inow, i = 0 and Q = Q0 , a constant, throughout. Dividing both sides of equation (2.45) by gA gives d S P/B S P/B = = F2 , (2.46) 2 dx 1 F 1/F 2 where, unusually for lectures on ow with a free surface, it has taken us until now to dene the Froude number F2 = Q2 B (Q/A) = , gA3 g (A/B)
2
the ratio of the mean velocity Q/A squared to g times the mean depth A/B. In this course we call F 2 the Froude number, and not F , as the latter quantity occurs quite rarely, and F 2 expresses the real relative importance of inertia terms. 31
The Gradually-Varied Flow Equation (GVFE) in the form of equation (2.46) seems simple deceptively simple. For example, can be taken as a constant; S might be a function of x, but we probably do not have enough information to express it as a function of ; many open channels are much wider than their depth, and so P B and P/B 1. This leaves most of the functional variation with on the right side in the term 1/F 2 = gA3 /Q2 B in which, for practical river problems the dependence of A and B on the local elevation is actually quite complicated. Constant slope: As a special case, consider a channel with a bed of constant slope S = S0 . It is simpler to use as a variable the depth of ow h, where h = Z, where Z is the elevation of the bed at a section, so that dZ/dx = S0 . Equation (2.46) becomes d S0 P/B dh dZ dh = + = S0 = . dx dx dx dx 1/F 2 Solving for dh/dx and introducing the conveyance gives the GVFE for a prismatic canal of constant slope: S0 Q2 /K 2 dh . = dx 1 F 2 Properties of gradually-varied ow and the governing equations (2.47)
Figure 2.13: Subcritical ow retarded by a gate, showing typical behaviour of the free surface and, if the channel is prismatic, decaying upstream to normal depth
The equation and its solutions are important, in that they tell us how far the effects of a structure or works in or on a stream extend upstream or downstream. It is an ordinary differential equation of rst order, hence one boundary condition must be supplied to obtain the solution. In sub-critical ow, this is the depth at a downstream control; in super-critical ow it is the depth at an upstream control. If the channel is prismatic, far from the control, the ow is uniform, and the depth is said to be normal. In general the boundary depth is not equal to the normal depth, and the differential equation describes the transition from the one to the other. The solutions look like exponential decay curves, and below we will show that they are, to a rst approximation. Figure 2.13 shows a typical sub-critical ow in a prismatic channel, where the depth at a control is greater than the normal depth. The differential equation is nonlinear, and the dependence on h is complicated, such that analytical solution is not possible, and we will usually use numerical methods. However, a small-disturbance approximation can be made, the resulting analytical solution is useful in providing us with some insight into the quantities which govern the extent of the upstream or downstream inuence. If the ow approaches critical ow, when F 2 1, then dh/dx , and the surface becomes vertical. This violates the assumption we made that the ow is gradually varied and the pressure distribution is hydrostatic. This is the one great failure of our open channel hydraulics at this level, that it cannot describe the transition between sub- and super-critical ow. Approximate analytical solution by linearising Whereas the numerical solutions give us numbers to analyse, sometimes very few actual numbers are required, such as merely estimating how far upstream water levels are raised to a certain level, the effect of downstream works on ooding, for example. Here we introduce a different way of looking at a physical problem in hydraulics, 32
where we obtain an approximate mathematical solution so that we can provide equations which reveal to us more of the nature of the problem than do numbers. This work was originally done by Samuels (1989). This is carried out by linearising about the uniform ow in a prismatic channel, i.e. by considering small disturbances to that ow. Consider the water depth to be written h(x) = h0 + h1 (x), (2.48)
where we use the symbol h0 for the constant normal depth, and h1 (x) is a relatively small departure of the surface from that. We use the governing differential equation (2.47) (although our notation has obscured the fact that F and K are functions of h). Substitute equation (2.48) into the equation and writing numerator and denominator as Taylor series in h1 : d S0 Q2 /K 2 0 + h1 dh S0 Q2 /K 2 0 + terms in h2 dh0 dh1 1 d . + = 2 2) + h 2) dx dx (1 F 0 + terms in h1 1 dh (1 F 0 Now, as h0 is constant, dh0 /dx = 0. Also, from equation (2.40), S0 Q2 /K 2 0 = 0 and the rst term in the numerator is zero. Now evaluating d S0 Q2 /K 2 /dh = 2Q2 /K 3 dK/dh, and considering just the rst term top and bottom, neglecting all higher order powers of h1 as it is small, we nd
3 2Q2 /K0 (dK/dh)0 dh1 = 0 h1 , h1 2 dx 1 F0
(2.49)
where 0 =
2 and where we have used Q2 /K0 = S0 .
2S0 1 dK , 2 1 F0 K dh 0
(2.50)
Equation (2.49) is a linear differential equation which we can solve analytically by separation of variables, giving h1 = Ce0 x , and h = h0 + Ce0 x , (2.51)
where C is a constant which would be evaluated by satisfying the boundary condition at the control, and where 0 is a constant decay rate given by equation (2.50). This shows that the water surface is actually approximated by an exponential curve passing from the value of depth 2 at the control to normal depth. As dK/dh is positive, and for subcritical ow 1 F0 is also positive, equation (2.50) shows that 0 is positive, and far upstream as x , the water surface decays to normal depth. For 2 supercritical ow, 1 F0 < 0, 0 is negative, and the water surface approaches normal depth downstream. Now we obtain an approximate expression for the rate of decay 0 . From the Gauckler-Manning-Strickler formula for a wide channel, a common approximation, we can show that K h5/3 , dK/dh 5/3 h2/3 , and for slow 2 ow F0 1, we nd 10 S0 0 . (2.52) 3 h0 The larger this number, the more rapid is the decay with x. The formula shows that more rapid decay occurs with steeper slopes (large S0 ), and smaller depths (h0 ). Hence, generally the water surface approaches normal depth more quickly for steeper and shallower streams, and the effects of a disturbance can extend a long way upstream for mild slopes and deeper water. Let us use equation (2.52) to calculate the distance upstream that the disturbance decays by 1/2, that is, exp (0 x) = 0.5. We nd x 3 ln 0.5 1 10 S0 x = ln 0.5 giving = . 3 h0 h0 10 S0 For S0 = 104 and a stream 2 m deep, the distance is 4 km. For the stream disturbance to decay to 1/16 = (1/2)4 of the original, this distance is 4 4 km = 16 km. These are possibly surprising results, showing how far the backwater effect extends. Numerical solution of the gradually-varied ow equation Consider the gradually-varied ow equation (2.47) dh S0 Q2 /K 2 = dx 1 F 2 33
where F 2 (h) = Q2 B(h)/gA3 (h). The equation is a differential equation of rst order, and to obtain solutions it is necessary to have a boundary condition h = h0 at a certain x = x0 , which will be provided by a control. The problem may be solved using any of a number of methods available for solving ordinary differential equations which are described in books on numerical methods. The direction of solution is very important. For mild slope (sub-critical ow) cases the surface decays somewhat exponentially to normal depth upstream from a downstream control, whereas for steep slope (super-critical ow) cases the surface decays exponentially to normal depth downstream from an upstream control. This means that to obtain numerical solutions we will always solve (a) for sub-critical ow: from the control upstream, and (b) for super-critical ow: from the control downstream. Eulers method The simplest (Euler) scheme to advance the solution from (xi , hi ) to (xi + xi , hi+1 ) is xi+1 hi+1 where xi is negative for subcritical ow, 2 2 dh = hi + xi S0 Q /K (hi ) + O (xi )2 . hi + xi 2 (h ) dx i 1 F i xi + xi , (2.53a) (2.53b)
This is the simplest but least accurate of all methods yet it might be appropriate for open channel problems where quantities may only be known approximately. One can use simple modications such as Heuns method to gain better accuracy, or use Richardson extrapolation or even more simply, just take smaller steps xi . Richardson extrapolation There is an interesting method for obtaining more accurate solutions from computational schemes for almost any physical problem. Applied to the Euler scheme in the present context, as the local truncation error is O (xi )2 , so that after a number of steps proportional to 1/xi , the actual solu 1 tion has an error O (xi ) so that n = 1, a rst-order scheme. In the present problem, if we use a constant space step to obtain the rst solution 1, then another constant space step half that /2, requiring twice the number of steps, then r = 1/2, and it can be shown that a better estimate of the solution is hi 2h2i (/2) hi (), (2.54)
which is trivially applied to each or any of the steps. The notation h2i (/2) is intended to show that the same point in physical space is used; with half the step size it will now take twice the number of steps to reach that point. Heuns method In this case the value of hi+1 calculated from Eulers method, equation (2.53b), is used as a rst estimate of the depth at the next point, written h , then the value of the derivative at that point xi+1 , h i+1 i+1 is calculated. Heuns method is then to use the mean slope over the step, the mean of the initial value and that at the other end of the interval calculated by the Euler step. Then, the change over the step is calculated, multiplying that mean slope by the step length. That is, hi+1 ! dh dh + dx (xi ,hi ) dx (xi+1 ,h ) i+1 2 2 xi S0 Q /K (hi ) S0 Q2 /K 2 (h ) 3 i+1 = hi + + + O (xi ) . 2 (h ) 2 (h ) 2 1 F 1 F i i+1 xi hi + 2
(2.55)
Now, the error of a single step is proportional to the third power of the step length and the error at any point will be proportional to the second power. Neither of these two methods are presented in hydraulics textbooks as alternatives, yet they are simple and exible, and reveal the nature of what we are doing. The step xi can be varied at will, to suit possible irregularly spaced cross-sectional data. In many situations, where F 2 1, we can ignore the F 2 term in the denominators, giving a notationally simpler scheme. Predictor-corrector method Trapezoidal method This is simply an iteration of the last method, whereby the step in equation (2.55) is repeated several times, at each stage setting h equal to the updated value of i+1 hi+1 . This gives an accurate and convenient method, and it is surprising that it has not been used. Higher-order methods One of the aims here has been to emphasise that all that is being done is to solve numerically a differential equation, and any method can be used, for which reference can be made to any book on numerical solution of ordinary differential equations. There are sophisticated methods such as high-order Runge-Kutta methods and predictor-corrector methods. However, in the case of open channel hydraulics there will usually be some variation of parameters along the channel that such sophistication is unnecessary. 34
2.6
2.4
2.2
Accurate: 4th order Runge-Kutta RK-4, n in error by 5% Euler, eqn (2.53) Euler-Richardson, eqns (2.53) & (2.54) Direct step, eqn (2.57) Modied direct step, eqn (2.58) Heun, eqn (2.55) Trapezoidal, eqn (2.55)+
1.8
Figure 2.14: Backwater curve computed with various schemes; the dashed line is the surface for uniform ow.
0.01
-0.02
-0.03
Figure 2.15:
-800
-600
x (m)
-400
-200
35
Comparison of schemes To compare the performance of the various numerical schemes, Example 10-1 of Chow (1959, p250) was solved using each. All quantities specied by Chow were converted to SI units and rounded to the numbers shown here: a ow of 11.33 m3 s1 passes down a trapezoidal channel of gradient S0 = 0.0016, bed width 6.10 m and channel side slopes 0.5, g = 9.8 m s2 , the quantity or = 1.10, and Mannings n = 0.025. At x = 0 the ow is backed up to a depth of 1.524 m, and the backwater curve was computed for 1000 m. Results for the water surface prole are shown in Figure 2.14, while Figure 2.15 shows the errors. Some 10 computational steps were used for each scheme. The basis of accuracy is shown by the solid line, from a highly-accurate Runge-Kutta 4th order method (see, for example, p372 of Yakowitz & Szidarovszky, 1989, or almost any other book on numerical methods). It is not recommended here as a method, however, as it makes use of information from three intermediate points at each step, information which in non-prismatic channels is usually not available. The rest of the results are shown in reverse order of accuracy. The dotted line is that with the same numerical method, but where the roughness n was changed by -5%, to give an idea of the effect of uncertainty in knowledge of that quantity. The maximum error, of about 3 cm, in the normal depth, is greater than any of the other methods, so that a preliminary conclusion is that if the roughness is not known to within 5%, almost certainly the case in practice, then any of the methods can be used. It can be seen that Eulers method, eqn (2.53), was the least accurate, as expected. As it is a rst-order scheme, halving the step size would halve the error. Actually doing just that and then applying Richardson extrapolation, equation (2.54), gave the second most accurate of all the methods, with an error of about 1 mm. The most accurate of all was the Trapezoidal method, namely using Heuns method, equation (2.55) and iterating the nal step. All the other methods, Heuns, and the two inverse formulations, equations (2.57) and (2.58), gave errors intermediate between the two extremes. It is interesting that the two traditional methods were accurate, notably the traditional inverse formulation over the modied version presented in this work; and the Trapezoidal method, the basis for the so-called "standard" method. The results show the disadvantages of the inverse formulation (Direct Step), that the distance between computational points becomes large as uniform ow is approached, and the points are at awkward distances. In this example relatively few steps were chosen (roughly 10) so that the numerical accuracies of the methods could be distinguished visually. The computational effort was very small indeed. In this problem the analytical solution (2.51) gave poor results. This was because the depth at the control was rather larger (50%) than the normal depth, and the linearisation adopted, for small departures from normal depth, was not accurate. In general, however, it does give a simple approximate result for the rate of decay and how far upstream the effects of the control extend. For many practical problems, this accuracy and simplicity may be enough. Traditional methods Here we present methods for comparison as they are given in textbooks.
2 U2 /2g
Sf x
2 U1 /2g
h2 S0 x
h1
x 2
Figure 2.16:
1
Elemental section of waterway
Derivation of the gradually-varied ow equation using energy Consider the elemental section of waterway of length x shown in Figure 2.16. We have shown stations 1 and 2 in what might be considered the reverse order, but for the more common sub-critical ow, numerical solution of the governing equation will proceed
36
back up the stream. Considering stations 1 and 2: Total head at 2 = Total head at 1 = H2 H1 = H2 HL ,
and we introduce the concept of the friction slope Sf which is the gradient of the total energy line such that HL = Sf x. This gives H1 = H2 Sf x, and if we introduce the Taylor series expansion for H1 : H1 = H2 + x substituting and taking the limit x 0 gives dH (2.56) = Sf , dx an ordinary differential equation for the head as a function of x, and we use the approximation that the friction slope is given by Q2 /K 2 . Direct step method Textbooks do present the Direct Step method, which is applied by taking steps in the height and calculating the corresponding step in x. It has practical disadvantages, such that it is applicable only to prismatic sections, results are not obtained at specied points in x, and as uniform ow is approached the x become innitely large. However it is a surprisingly accurate method. The reciprocal of equation (2.56) is dx 1 = . dH Sf The numerical method as set out in textbooks is to approximate the differential equation (2.56) by the nite difference expression xi = = Hbi S0 Q2 /K 2 (Hb ) Hbi 2 2 S0 1 Q2 Ki + Ki+1 2 (2.57a) (2.57b) dH + ..., dx
The term O (. . .) is a Landau order symbol, showing in this case that the local truncation error is proportional to the third power of the step, which is a strong result and explains the accuracy of the method. Since the use of a step size of Hb,i over the whole computational domain requires a number of steps proportional to 1/Hb,i , 2 the global error in this case will be of order (Hb,i ) , thus the global error, or accumulated error at the end of that integration interval will be of this order, so that halving the step should improve the global accuracy by about a factor of 4. In view of the method presented here, the method is no longer applicable only to prismatic sections, but the practical disadvantages remain that results are not obtained at specied points in x, and as uniform ow is approached the x become innitely large. Standard step method The nomenclature "standard" is not very descriptive. Presumably it refers to nding the solution for at specied values of x, rather than the other way round, for which the term "direct", as above, is even worse. This is an implicit method, requiring numerical solution of a transcendental equation at each step. It can be used for irregular channels, and is rather more general. In this case, the distance interval x is specied and the corresponding depth change calculated. In the Standard step method the procedure is to write H = Sf x, 37
where the overbar in equation (2.57a) indicates the mean of the friction slope at beginning and end of the computational interval, which nds its mathematical expression in equation (2.57b), where the shorthand Ki has been used for K (Hbi ). While this is a plausible approximation, it is not mathematically consistent. It is an apparent attempt to develop a Trapezoidal method. Applying Heuns method as formally presented in equation (2.55) automatically leads to the Trapezoidal scheme which in this case gives Hb,i 1 1 xi+1 = xi + + + O (Hb,i )3 , (2.58) 2 /K 2 2 /K 2 2 S0 Q S0 Q i i+1
x (Sf1 + Sf2 ) , 2 for sections 1 and 2, where the mean value of the friction slope is used. This gives H2 (h2 ) H1 (h1 ) = Q2 Q2 x + Z2 + h2 = 2 2 + Z1 + h1 2 (Sf1 + Sf2 ) , 2gA2 2gA1
where Z1 and Z2 are the elevations of the bed. This is a transcendental equation for h2 , as this determines A2 , P2 , and Sf2 . Solution could be by any of the methods we have had for solving transcendental equations, such as direct iteration, bisection, or Newtons method. Although the Standard step method is an accurate and stable approximation, the lecturer considers it unnecessarily complicated, as it requires solution of a transcendental equation at each step. It would be much simpler to use a simple explicit Euler or Heuns method as described above.
It is possible to simplify the usual approach of considering two long wave equations in two unknowns by introducing a quantity that satises one of the equations exactly. We consider the (A, Q) formulation of the long wave equations, equations (2.22). In partial derivative terms we introduce the function V (x, t) which is the total volume of uid upstream of point x at time t, and use equations (2.5) in 2.3.1.4 so that we write: Z x V V A= i dx0 Q, and = x t and in the latter quantity the rate at which upstream volume is increasing in time is the integral of inow minus the ow which is currently actually passing point x. Solving the latter for Q and substituting into the mass conservation equation (2.22a) Z x V V 0 i dx + = i, t x x t and the mass conservation equation is identically satised!The momentum conservation equation (2.22b), with all signs reversed becomes: 2 R x Z i dx0 V /t 2V Q 2V gA 2 V Q2 + + 2 = P gAS + 2 2 t2 A xt A B x2 (V /x)
x
i 0 Q dx + 2 i. (2.59) t A
where for notational simplicity, single symbols Q and A have been retained in coefcients of derivatives. The momentum equation has become a second-order partial differential equation in terms of the single variable V . This makes theoretical manipulations easier and provides insight into the nature of solutions of the equation. This is anticipated by the comment that the rst three terms on the left actually look like a wave equation the traditional forms of equations (2.22) and (2.23) do not look like wave equations. However the single equation certainly does look complicated it will be of most use when we consider approximations. The case of no inow i = 0 is rather simpler, and is adequate for our purposes. Equation (2.59) becomes 2V Q 2V gA 2 V (V /t)2 Q2 + 2 = P + 2 2 gAS. t2 A xt A B x2 (V /x) (2.60)
This equation, which is a single equation in a single unknown and could be used for ood routing and wave propagation studies. Here we will use it rst to obtain an approximate equation which is then solved to give solutions that show the real behaviour of long waves.
2.4.2
Derivation of equation As in the analytical solution for steady ow of 2.3.6.3, here we linearise the equation by considering small perturbations about a uniform ow of cross-sectional area A0 and discharge Q0 in a channel of constant slope, 38