Vous êtes sur la page 1sur 70

November 14, 2008

River hydraulics
John Fenton

Abstract
This elective describes the nature of ows in rivers, their measurement, the calculation of ows and ood propagation, the knowledge of the factors affecting water quality, and the stability of rivers. We will see how some of the fundamentals have been glossed over, and some very simple improvements can be made to traditional practice. It is hoped that students taking this course will develop a deep understanding of the processes at work in rivers. Throughout these lectures, both in approximations to wave motion in waterways, and in the transport of pollutants, we will encounter the physical process of diffusion. An introduction to diffusion is given in Appendix A.1, but is not for examination.

Table of Contents
References 1. 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 8

Introduction .

Hydrography/Hydrometry . . . . . . . . . . . . 2.1 Water levels . . . . . . . . . . . . . . . 2.2 Discharge . . . . . . . . . . . . . . . 2.3 The analysis and use of stage and discharge measurements

. 8 . 8 . 9 . 16 . . . . . . . . . . . . . . . 24 25 26 29 33 37 37 41 42 43 43 43 43 44 44

3.

The propagation of waves in waterways . . . . . . . . . 3.1 Mass conservation equation . . . . . . . . . . . 3.2 Momentum conservation equation . . . . . . . . . 3.3 The nature of the propagation of long waves and oods in rivers 3.4 A new low-inertia approach Volume routing . . . . . Computational hydraulics . . . 4.1 The advection equation . . 4.2 The diffusion equation . . 4.3 Advection-diffusion combined . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.

5.

Water quality . . . . . . . . 5.1 Useful sources for further reading 5.2 Water quality characteristics . . 5.3 Types of pollutant . . . . . 5.4 Mass balance concepts . . . 5.5 Impacts of human works . . . 1

River hydraulics

John Fenton

5.6 5.7 5.8 5.9 6.

Transport processes . . . . . . . . . . . . Tools for problem solving . . . . . . . . . . A simple river model organic wastes and self purication Salinity in rivers . . . . . . . . . . . . .

. . . .

. . . .

. . . . . . . . . . . . . .

. . . .

45 46 46 53

Turbulent diffusion and dispersion . . . . . . . . . . . . 6.1 Diffusion and dispersion in waterways . . . . . . . . . 6.2 Dispersion . . . . . . . . . . . . . . . . . 6.3 Non-dimensionalisation Pclet number and Reynolds number viscosity as diffusion . . . . . . . . . . . . . . Sediment motion . . . . . . . 7.1 Incipient motion . . . . . 7.2 Relationships for uvial quantities 7.3 Dimensional similitude . . . 7.4 Bed forms . . . . . . . 7.5 Mechanisms of sediment motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. 56 . 57 . 57 . 59 . . . . . . 59 60 61 62 62 63

7.

Appendix A On diffusion and von Neumann stability analyses . . . . . . A.1 The nature of diffusion . . . . . . . . . . . . . . A.2 Examining stability by the Fourier series (von Neumanns) method .

. 65 . 65 . 69

River hydraulics

John Fenton

Useful references
Tables 1.1-1.4 show some of the many references available, some which the lecturer has referred to in these notes or in his work.
Reference 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. & 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. 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. Comments Good technical book, moderate level, also sediment aspects Good technical book Classic, now dated, not so readable Good elementary introduction Wide general treatment Classic, high level, readable High level, but terse and readable A readable but high-level work Encyclopaedic Simple, readable, mathematical Simple introduction to computational hydraulics

Table 1.1 : Introductory and general references

Reference Australian Standard 3778 (1990), Australian Standard - Measurement of water ow in open channels, Standards Association of Australia, Homebush. 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, Co-operative Research Centre for Catchment Hydrology, Monash University. Novak, P., Moffat, A. I. B., Nalluri, C. & Narayanan, R. (2001), Hydraulic Structures, third edn, Spon, London.

Comments Could be much better than it is 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

Table 1.2 : Books on practical aspects, ow measurement, and structures

River hydraulics

John Fenton

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. 205230. Fenton, J. D. (1996), An examination of the approximations in river and channel hydraulics, in Proc. 10th Congress, Asia-Pacic Division, Int. Assoc. Hydraulic Res., Langkawi, Malaysia, pp. 204211. 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 numer ical schemes modelling open channel ow, in C. Maksimovi & M. Radojkovi , c c eds, Computational Modelling and Experimental Methods in Hydraulics (HYDROCOMP 89), Elsevier. Zoppou, C. & ONeill, I. C. (1982), Criteria for the choice of ood routing methods in natural channels, in Proc. Hydrology and Water Resources Symposium, Melbourne, pp. 7581.

Comments Thorough and reliable presentation Interesting review

A modern mathematical view

A readable overview Good review, a bit dated Readable overview

Readable overview

Readable

The best overview of the advectiondiffusion approximation for ood routing Review

Readable overview

Table 1.3 : References on ood wave propagation theoretical and computational

River hydraulics

John Fenton

Reference General Chin (2000) Martin & McCutcheon (1999)

Notes A good introduction A good book, being both introductory and encyclopaedic, concentrating on the hydraulic engineering aspects A complete descriptive (non-mathematical) presentation, which is interesting. A comprehensive and standard reference Also fundamental, but shorter A good simpler introduction (Chapter 9)

McGauhey (1968) Fundamental processes of mixing and dispersion Fischer, List, Koh, Imberger & Brooks (1979) Holly (1985) Streeter, Wylie & Bedford (1998) Rutherford (1994) Csanady (1973) Numerical methods fundamentals Noye (1976), Noye (1981), Noye (1984), Noye & May (1986)

All offer a simple introduction to nite difference methods Smith (1978) A more detailed introduction to nite difference methods Richtmyer & Morton (1967), Morton & Baines (1982), Morton & Mayers (1994), Morton All are rather more comprehensive, describ(1996) ing some more general methods Zoppou & Knight (1997) Analytical solutions to the advectiondiffusion equation where the coefcients are not constant Numerical methods application to environmental modelling Sauvaget (1985) A simple review The nature of diffusion Fischer et al. (1979) Very clear - already recommended above Jost (1960, page 25; 1964) A leisurely and clear introduction Borg & Dienes (1988) A simple and clear introduction Widder (1975) A more mathematical approach The full equations for wave propagation and ood routing Cunge, Holly & Verwey (1980) The best explanation of this eld Liggett (1975), Liggett & Cunge (1975) A little disappointing, but the next best explanation The advection-diffusion approximation for ood routing Price (1985) The best overview Dooge (1986) A good general study Sivapalan, Bates & Larsen (1997) Others Pasmanter (1988) Estuaries and tidal ows Kobus & Winzelbach (1989) Groundwater

Table 1.4 : Useful references

River hydraulics

John Fenton

References
Australian Standard (1990) Measurement of water ow in open channels, number AS 3778, Standards Australia. Australian Standard 3778.3.1 (2001) Measurement of water ow in open channels - Velocity-area methods - Measurement by current meters and oats, Standards Australia, Sydney. Boiten, W. (2000) Hydrometry, Balkema. Borg, R. J. & Dienes, G. J. (1988) An Introduction to Solid State Diffusion, Academic. Chin, D. A. (2000) Water-Resources Engineering, Prentice Hall. Chow, V. T. (1959) Open-channel Hydraulics, McGraw-Hill, New York. Collett, K. O. (1978) The present salinity position in the River Murray basin, Proc. Royal Society of Victoria 90(1), 111123. Csanady, G. T. (1973) Turbulent Diffusion in the Environment, Reidel, Dordrecht. Cunge, J. A., Holly, F. M. & Verwey, A. (1980) Practical Aspects of Computational River Hydraulics, Pitman, London. Dooge, J. C. I. (1986) Theory of ood routing, River Flow Modelling and Forecasting, D. A. Kraijenhoff & J. R. Moll (eds), Reidel, chapter 3, pp. 3965. Elmore, H. L. & Hayes, T. W. (1960) Solubility of atmospheric oxygen in water, J.Sanitary Div. ASCE 86(SA4), 4153. Fenton, J. D. (1999) Calculating hydrographs from stage records, in Proc. 28th IAHR Congress, 22-27 August 1999, Graz, Austria, published as compact disk. Fenton, J. D. (2002) The application of numerical methods and mathematics to hydrography, in Proc. 11th Australasian Hydrographic Conference, Sydney, 3 July - 6 July 2002. Fenton, J. D. & Abbott, J. E. (1977) Initial movement of grains on a stream bed: the effect of relative protrusion, Proc. Roy. Soc. Lond. A 352, 523537. 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, Melbourne. http: //www.catchment.crc.org.au/pdfs/technical200106.pdf Feynman, R. P. (1985) Surely youre joking, Mr. Feynman! : adventures of a curious character, Norton, New York. Fischer, H. B., List, E. J., Koh, R. C. Y., Imberger, J. & Brooks, N. H. (1979) Mixing in Inland and Coastal Waters, Academic. French, R. H. (1985) Open-Channel Hydraulics, McGraw-Hill, New York. Goldsmith, E. & Hildyard, N. (1992) The Social and Environmental Effects of Large Dams, Wadebridge Ecological Centre, Camelford, Cornwall, UK. Henderson, F. M. (1966) Open Channel Flow, Macmillan, New York. Herschy, R. W. (1995) Streamow Measurement, Second Edn, Spon, London. Holly, F. M. (1985) Dispersion in rivers and coastal waters 1. Physical principles and dispersion equations, Developments in Hydraulic Engineering, P. Novak (ed.), Vol. 3, Elsevier, London, chapter 1. Jost, W. (1960) Diffusion in Solids, Liquids, Gases, Academic, New York. Jost, W. (1964) Fundamental Aspects of Diffusion Processes, Angewandte Chemie Int. Edn 3, 713722. Keiller, D. & Close, A. (1985) Modelling salt transport in a long river system, in Proc. 21st Congress IAHR, Melbourne, Vol. 2, pp. 324328. 6

River hydraulics

John Fenton

Kobus, H. E. & Winzelbach, W. (1989) Contaminant Transport in Groundwater, Balkema, Rotterdam. 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. Lighthill, M. J. & Whitham, G. B. (1955) On kinematic waves. I: Flood movement in long rivers, Proc. R. Soc. Lond. A 229, 281316. Martin, J. L. & McCutcheon, S. C. (1999) Hydrodynamics and Transport for Water Quality Modeling, Lewis, Boca Raton. McGauhey, P. H. (1968) Engineering Management of Water Quality, McGraw-Hill, New York. Morgan, A. E. (1971) Dams and other disasters: a History of the Army Corps of Engineers, Porter Sargent, Boston. Morton, K. (1996) Numerical solution of convection-diffusion problems, Chapman and Hall, London. Morton, K. & Baines, M. (1982) Numerical methods for uid dynamics, Academic. Morton, K. & Mayers, D. (1994) Numerical solution of partial differential equations : an introduction, Cambridge. Noye, B. J. (1976) International Conference on the Numerical Simulation of Fluid Dynamic Systems, Monash University 1976, North-Holland, Amsterdam. Noye, B. J. (1981) Numerical solutions to partial differential equations, Proc. Conf. on Numerical Solutions of Partial Differential 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 differential equations, North-Holland, Amsterdam. Noye, J. & May, R. L. (1986) Computational Techniques and Applications: CTAC 85, North-Holland, Amsterdam. Pasmanter, R. A. (1988) Deterministic diffusion, effective shear and patchiness in shallow tidal ows, Physical Processes in Estuaries, J. Dronkers & W. van Leussen (eds), Springer, Berlin. Price, R. K. (1985) Flood Routing, Developments in Hydraulic Engineering, P. Novak (ed.), Vol. 3, Elsevier Applied Science, chapter 4, pp. 129173. Richtmyer, R. P. & Morton, K. W. (1967) Difference Methods for Initial Value Problems, Second Edn, Interscience, New York. Rutherford, J. C. (1994) River Mixing, Wiley, Chichester. Sauvaget, P. (1985) Dispersion in rivers and coastal waters 2. Numerical computation of dispersion, Developments in Hydraulic Engineering, P. Novak (ed.), Vol. 3, Elsevier, London, chapter 2. Schlichting, H. (1968) Boundary-Layer Theory, Sixth Edn, McGraw-Hill, New York. Sivapalan, M., Bates, B. C. & Larsen, J. E. (1997) A generalized, non-linear, diffusion wave equation: theoretical development and application, Journal of Hydrology 192, 116. Smith, G. D. (1978) Numerical Solution of Partial Differential Equations, Oxford Applied Mathematics and Computing Series, Second Edn, Clarendon, Oxford. Streeter, V. L., Wylie, E. B. & Bedford, K. W. (1998) Fluid Mechanics, Ninth Edn, WCB/McGraw-Hill. Widder, D. V. (1975) The Heat Equation, Academic, New York. Yalin, M. S. & Ferreira da Silva, A. M. (2001) Fluvial Processes, IAHR, Delft. Zoppou, C. & Knight, J. H. (1997) Analytical solutions for advection and advection-diffusion equations with spatially variable coefcients, Journal of Hydraulic Engineering 123(2), 144148. 7

River hydraulics

John Fenton

1. Introduction
At the conclusion of this unit, students should be able to describe the nature of ow and oods in streams, understand the basis of computational methods for rivers, the common means of measurement of streamow, the fundamentals of water quality in rivers, and uvial processes and uvial morphology.

2. Hydrography/Hydrometry
Boiten (2000) provides a refreshingly modern approach to this topic, calling it Hydrometry the measurement of water, which in the past has received little research. In particular, the Australian Standard (1990) is a very poor document, providing little practical assistance.

2.1 Water levels


Water levels are the basis for any river study. Most kinds of measurements, such as discharges, have to be related to river stages (the stage is simply the water surface height above some xed datum). Both stage and discharge measurements are important. Often, however, the actual discharge of a river is measured rarely, and routine measurements are those of stage, which are related to discharge. Water levels are obtained from gauges, either by direct observation or in recorded form. The latter is now much more likely in the Australian water industry. The data can serve several purposes:
By plotting gauge readings against time, the hydrograph for a particular station is obtained. Hydrographs of a series of years are used to determine duration curves, showing the probability of occurrence of water levels at the station or from a rating curve, the probability of discharges. Combining gauge readings with discharge values, a relationship between stage and discharge can be determined, resulting in a rating curve for the station. Apart from use in hydrological studies and for design purposes, the data can be of direct value for navigation, ood prediction, water management, and waste water disposal.

2.1.1 Methods Most water level gauging stations are equipped with a sensor or gauge and a recorder. In many cases the water level is measured in a stilling well, thus eliminating strong oscillations. Staff gauge: This is the simplest type, with a graduated gauge plate xed to a stable structure such as a pile, bridge pier, or a wall. Where the range of water levels exceeds the capacity of a single gauge, additional ones may be placed on the line of the cross section normal to the plane of ow. Float gauge: A oat inside a stilling well, connected to the river by an inlet pipe, is moved up and down by the water level. Fluctuations caused by short waves are almost eliminated. The movement of the oat is transmitted by a wire passing over a oat wheel, which records the motion, leading down to a counterweight. Pressure transducers: The water level is measured as an equivalent hydrostatic pressure and transformed into an electrical signal via a semi-conductor sensor. These are best suited for measuring water levels in open water (the effect of short waves dies out almost completely within half a wavelength down into the water), as well as for the continuous recording of groundwater levels. They should compensate for changes in the atmospheric pressure, and if air-vented cables cannot be provided air pressure needs to be measured separately. Bubble gauge: This is a pressure actuated system, based on measurement of the pressure which is needed to produce bubbles through an underwater outlet. These are used at sites where it would be 8

River hydraulics

John Fenton

difcult to install a oat-operated recorder or pressure transducer. From a pressurised gas cylinder or small compressor gas is led along a tube to some point under the water (which will remain so for all water levels) and bubbles constantly ow out through the orice. The pressure in the measuring tube corresponds to that in the water above the orice. Wind waves should not affect this. Ultrasonic sensor: These are used for continuous non-contact level measurements in open channels, and are widely used in the Australian irrigation industry. The sensor points vertically down towards the water and emits ultrasonic pulses at a certain frequency. The inaudible sound waves are reected by the water surface and received by the sensor. The round trip time is measured electronically and appears as an output signal proportional to the level. A temperature probe compensates for variations in the speed of sound in air. They are accurate but susceptible to wind waves. Peak level indicators: There are some indicators of the maximum level reached by a ood, such as arrays of bottles which tip and ll when the water reaches them, or a staff coated with soluble paint. 2.1.2 Presentation of results Stage records taken along rivers used for hydrological studies, for design of irrigation works, or for ood protection require an accuracy of 2 5 cm, while gauge readings upstream of ow measuring weirs used to calculate discharges from the measured heads require an accuracy of 2 5 mm. These days almost all are telemetered to a central site. There is a huge volume of electronic hydrometry data being sent around Victoria. Hydrographs, rating tables, and stage relation curves are typical presentations of water level data:
Hydrograph when stage records or the discharges are plotted against time. Rating table at many gauging stations water levels are measured daily or hourly, while discharges are measured some times a year, using direct methods such as a propeller meter. From the corresponding water levels from these, and possibly for others over years, a stage-discharge relationship can be built up, so that the routine measurement of stage can be converted to discharge. We will be considering these in detail. Stage relation curves from the hydrographs of two or more gauging stations along the river, relationships can be formulated between the steady ow stages. These can be used to calculate the surface slope between two gauges, and hence, to determine the roughness of the reach. Under unsteady conditions the relationship will be disturbed. We will also be considering this later.

2.2 Discharge
Flow measurement may serve several purposes:
information on river ow for the design and operation of diversion dams and reservoirs and for bilateral agreements between states and countries. distribution and charging of irrigation water information for charging industries and treatment plants discharging into public waters water management in urban and rural areas reliable statistics for long-term monitoring.

Continuous (daily or hourly) measurements are very useful. There are many methods of measuring the rate of volume ow past a point, of which some are single measurement methods which are not designed for routine operation; the rest are methods of continuous measurements.

River hydraulics

John Fenton

2.2.1 Velocity area method (current meter method) The area of cross-section is determined from soundings, and ow velocities are measured using propeller current meters, electromagnetic sensors, or oats. The mean ow velocity is deduced from points distributed systematically over the river cross-section. In fact, what this usually means is that two or more velocity measurements are made on each of a number of vertical lines, and any one of several empirical expressions used to calculate the mean velocity on each vertical, the lot then being integrated across the channel. Calculating the discharge requires integrating the velocity data over the whole channel - what is required R is the area integral of the velocity, that is Q = u dA. If we express this as a double integral we can write
Q= Z
Z(y)+h(y) Z Z(y)

u dz dy,

(2.1)

so that we integrate the velocity from the bed z = 0 to the surface z = h(y), where h is the local depth and where our z is a local co-ordinate. Then we have to integrate these contributions right across the channel, for values of the transverse co-ordinate z over the breadth B . Calculation of mean velocity in the vertical The rst step is to compute the integral of velocity with depth, which hydrographers think of as calculating the mean velocity over the depth. Convention in hydrography is that the mean velocity over a vertical can be approximated by 1 u = (u0.2h + u0.8h ) , (2.2) 2 that is, the mean of the readings at 0.2 of the depth and 0.8 of the depth. Fenton (2002) has developed some families of methods which are based more on rational methods. Consider the law for turbulent ow over a rough bed, which can be obtained from the expressions on p582 of Schlichting (1968):
u= z u ln , z0

(2.3)

where u is the shear velocity, = 0.4, ln() is the natural logarithm to the base e, z is the elevation above the bed, and z0 is the elevation at which the velocity is zero. (It is a mathematical artifact that below this point the velocity is actually negative and indeed innite when z = 0 this does not usually matter in practice). If we integrate equation (2.3) over the depth h we obtain the expression for the mean velocity:
1 u= h Zh
0

u u dz =

h ln 1 . z0

(2.4)

Now it is assumed that two velocity readings are made, obtaining u1 at z1 and u2 at z2 . This gives enough information to obtain the two quantities u / and z0 . Substituting the values for point 1 into equation (2.3) gives us one equation and the values for point 2 gives us another equation. Both can be solved to give the solution u2 1 u2 u1 u z1 u2 u1 = and z0 = . (2.5) u ln (z2 /z1 ) z2 1 It is not necessary to evaluate these, for substituting into equation (2.4) gives a simple formula for the mean velocity in terms of the readings at the two points:
u= u1 (ln(z2 /h)+1) u2 (ln(z1 /h)+1) . ln (z2 /z1 )

(2.6)

As it is probably more convenient to measure and record depths rather than elevations above the bottom, 10

River hydraulics

John Fenton

let h1 = h z1 and h2 = h z2 be the depths of the two points, when equation (2.6) becomes
u= u1 (ln(1 h2 /h)+1) u2 (ln(1 h1 /h)+1) . ln ((h h2 ) / (h h1 ))

(2.7)

This expression gives the freedom to take the velocity readings at any two points, and not necessarily at points such as 0.2h and 0.8h. This might simplify streamgauging operations, for it means that the hydrographer, after measuring the depth h, does not have to calculate the values of 0.2h and 0.8h and then set the meter at those points. Instead, the meter can be set at any two points, within reason, the depth and the velocity simply recorded for each, and equation (2.7) applied. This could be done either in situ or later when the results are being processed. This has the potential to speed up hydrographic measurements. If the hydrographer were to use the traditional two points, then setting h1 = 0.2h and h2 = 0.8h in equation (2.7) gives the result
u = 0.4396 u0.2h + 0.5604 u0.8h 0.44u0.2h + 0.56u0.8h ,

(2.8)

whereas the conventional hydrographic expression is (see e.g. #7.1.5.3 of Australian Standard 3778.3.1 2001):
u = 0.5 u0.2h + 0.5 u0.8h .

(2.9)

The nominally more accurate expression, equation (2.8), gives less weight to the upper measurement and more to the lower. It might be useful, as it is just as simple as the traditional expression, yet is based on an exact analytical integration of the equation for a turbulent boundary layer. This has been tested by taking a set of gauging results. A canal had a maximum depth of 2.6m and was 28m wide, and a number of verticals were used. The conventional formula (2.2), the mean of the two velocities, was accurate to within 2% of equation (2.8) over the whole range of the readings, with a mean difference of 1%. That error was always an overestimate. The more accurate formula (2.7) is hardly more complicated than the traditional one, and it should in general be preferred. Although the gain in accuracy was slight in this example, in principle it is desirable to use an expression which makes no numerical approximations to that which it is purporting to evaluate. This does not necessarily mean that either (2.2) or (2.8) gives an accurate integration of the velocities which were encountered in the eld. In fact, one complication is where, as often happens in practice, the velocity distribution near the surface actually bends back such that the maximum velocity is below the surface. Fenton (2002) then considered velocity distributions given by the more general law, assuming an additional linear and an additional quadratic term in the velocity prole:
u= z u ln + a1 z + a2 z 2 , z0

(2.10)

and by taking readings at four depths, enough information is obtained to obtain the solution for u. Methods and computer code for this were presented. Also, in something of a departure, a global approximation method was used, where a function was assumed which could describe all the velocity proles on all the verticals, and then this was tted to the data. An example of the results is given

Figure 2-1. Cross-section of canal with velocity proles and data points plotted transversely, showing t by global function

11

River hydraulics

John Fenton

Integration of the mean velocities across the channel: The problem now is to integrate the readings for mean velocity at each station across the width of the channel. Here traditional practice seems to be in error often the Mean-Section method is used. In this the mean velocity between two verticals is calculated and then multiply this by the area between them, so that, given two verticals i and i + 1 separated by bi the expression for the contribution to discharge is assumed to be 1 Qi = bi (hi + hi+1 ) (ui + ui+1 ) . 4 This is not correct. From equation (2.1), the task is actually to integrate across the channel the quantity which is the mean velocity times the depth. For that the simplest expression is the Trapezoidal rule:
1 Qi = bi (ui+1 hi+1 +ui hi ) 2

To examine where the Mean-Section Method is worst, we consider the case at one side of the channel, where the area is a triangle. We let the waters edge be i = 0 and the rst internal point be i = 1, then the Mean-Section Method gives 1 Q0 = b0 u1 h1 , 4 while the Trapezoidal rule gives 1 Q0 = b0 u1 h1 , 2 which is correct, and we see that the Mean-Section Method computes only half of the actual contribution. The same happens at the other side. Contributions at these edges are not large, and in the middle of the channel the formula is not so much in error, but in principle the Mean-Section Method is wrong and should not be used. Rather, the Trapezoidal rule should be used, which is just as easily implemented. In a gauging in which the lecturer participated, a ow of 1693 Ml/d was calculated using the Mean-Section Method. Using the Trapezoidal rule, the ow calculated was 1721 Ml/d, a difference of 1.6%. Although the difference was not great, practitioners should be discouraged from using a formula which is wrong. In fact the story is rather more scandalous, because at least one ultrasonic method uses the Mean-Section Method for integrating vertically over only three or four data points, when its errors would be rather larger. In textbooks one does nd an approximate method known as the Mid-Section Method, which takes as the elemental contribution 1 Qi = ui hi (bi + bi+1 ) . 2 When the individual contributions are summed this becomes the Trapezoidal Rule. 2.2.2 Slope area method This is widely used to calculate peak discharges after the passage of a ood. An ideal site is a reach of uniform channel in which the ood peak prole is dened on both banks by high water marks. From this information the slope, the cross-sectional area and wetted perimeter can be obtained, and the discharge computed with the Gauckler-Manning-Strickler (G-M-S) formula or the Chzy formula. To do this however, roughness coefcients must be known, such as Mannings n in the G-M-S formula
1 A5/3 p S, n P 2/3 where A is the area, P the wetted perimeter, and S the slope. Q=

2.2.3 Dilution methods In channels where cross-sectional areas are difcult to determine (e.g. steep mountain streams) or where 12

River hydraulics

John Fenton

ow velocities are too high to be measured by current meters dilution or tracer methods can be used, where continuity of the tracer material is used with steady ow. The rate of input of tracer is measured, and downstream, after total mixing, the concentration is measured. The discharge in the stream immediately follows. 2.2.4 Integrating oat methods There is another rather charming and wonderful method which has been very little exploited. At the moment it has the status of a single measurement method, however the lecturer can foresee it being developed as a continuing method. Theory Consider a single buoyant particle (a oat, an orange, an air bubble), which is released from a point on the bed. We assume that it has a constant rise velocity w. As it rises it passes through a variable horizontal velocity eld u(z), where z is the vertical co-ordinate. The kinematic equations of the oat are
dx dt dz dt = u(z), = w.

Dividing the left and right sides, we obtain a differential equation for the particle trajectory
w dz = , dx u(z)

however this can be re-arranged as:


h Z 0

u(z) dz =

L Z 0

w dx = wL,

where the particle reaches the surface a distance L downstream of the point at which it was released on the bed, and where we have used a local vertical co-ordinate z with origin on the bed and where the uid locally has a depth h. The quantity on the left is important - it is the vertical integral of the horizontal velocity, or the discharge per unit width at that section. We can generalise the expression for variation with y, across the channel, to write
h Z

u(y, z) dz = wL(y),

Z(y)

where Z(y) is the z co-ordinate of the bed. Now, if we integrate across the channel, in the co-ordinate direction y , the integral of the left side is the discharge Q:
Q=
B h Z Z

u(y, z) dz dy = w

0 Z(y)

B Z 0

L(y) dy,

where B is the total width of the channel. Hence we have an expression for the discharge with very few approximations:
Q=w
B Z 0

L(y) dy.

13

River hydraulics

John Fenton

Figure 2-2. Array of four ultrasonic beams in a channel

If we were to release bubbles from a pipe across the bed of the stream, on the bed, then this is
Q = Bubble rise velocity area on surface between bubble path pattern and line of release.

This is possibly the most direct and potentially the most accurate of all ow measurement methods! 2.2.5 Ultrasonic ow measurement This is a method used in the irrigation industry in Australia, but is also being used in rivers in the USA. Consider the situation shown in the gure, where some three or four beams of ultrasonic sound are propagated diagonally across a stream at different levels. The time of travel of sound in one direction is measured, as is the time in the other. The difference can be used to compute the mean velocity along that path, i.e. at that level. These values then have to be integrated in the vertical. Mean velocity along beam path Unfortunately, in all textbooks and International and Australian Standards a constant velocity is assumed - precisely what is being sought to measure, and totally ignoring the fact that velocity varies along the path and indeed is zero at the ends! Here we include the variability of velocity in our analysis. Consider a velocity vector inclined to the beam path at an angle . If the velocity is u(s), showing that the velocity does, in general, depend on position along the beam, then the component along the path is u(s) cos . Let c be the speed of sound. The time dt taken for a sound wave to travel a distance ds along the path against the general direction of ow is dt = ds/ (c u(s) cos ). If the path has total length L, then the total time of travel T1 is obtained by integrating to give
T1 =
T Z1 0

dt =

ZL
0

ds , c u(s) cos

(2.11)

and repeating for a traverse in the reverse direction:


T2 =
T Z2 0

dt =

ZL
0

ds . c + u(s) cos

(2.12)

Now we expand the denominators of both integrals by the binomial theorem:


1 T1 = c ZL ZL 1 u(s) u(s) cos ds and T2 = cos ds, 1+ 1 c c c
0 0

(2.13)

where we have ignored terms which contain the square of the uid velocity compared with the speed of

14

River hydraulics

John Fenton

sound. Evaluating gives


1 L T1 = + 2 c c ZL
0

1 L u(s) cos ds and T2 = 2 c c

ZL
0

u(s) cos ds.

(2.14)

Adding the two equations and solving for c and re-substituting we obtain
ZL
0

u(s) cos ds = 2L2

T1 T2 . (T1 + T2 )2

(2.15)

u It can be shown that the relative error of this expression is of order (/c)2 , where u is a measure of velocity. As u 1 m s1 and c 1400 m s1 it can be seen that the error is exceedingly small. What we rst need to compute the ow is the integral of the velocity component transverse to the beam path, for which we use the symbol Qz , the symbol with subscript suggesting the derivative of discharge with respect to elevation: Z L Qz = u(s) sin ds. (2.16)
0

Now we are forced to assume that the angle that the velocity vector makes with the beam is constant over the path (or at least in some rough averaged sense), and so for constant, taking the trigonometric functions outside the integral signs and combining equations (2.15) and (2.16) we obtain
Qz = 2 tan L2 T1 T2 . (T1 + T2 )2

(2.17)

This shows how the result is obtained by assuming the angle of inclination of the uid velocity to the beam is constant, but importantly it shows that it is not necessary to assume that velocity u is constant over the beam path. Equation (2.17) is similar to that presented in Standards and trade brochures, and implemented in practice, but where it is obtained by assuming that the velocity is constant. It is fortunate that the end result is correct. Vertical integration of beam data The mean velocities on different levels obtained from the beam data are considered to be highly accurate, provided all the technical problems associated with beam focussing etc. are overcome, and the streamow has a constant angle to the beam. The problem remains to calculate the discharge in the channel by evaluating the vertical integral of Qz , which, as shown by equation (2.16), is the integral along the beam of the velocity transverse to the beam. The problem is then to evaluate the vertical integral of the derivative of discharge with elevation:
Q=
h Z 0

Qz (z) dz,

(2.18)

where in practice the information available is that Qz = 0 on the bottom of the channel z = 0 and the two to four values of Qz which have been obtained from beam data, as well as the total depth h. It is in the evaluation of this integral that the performance of the trade and scientic literature has been poor. Several trade brochures advocate the routine use of a single beam, or maybe two, suggesting that that is adequate (see, for example, Boiten 2000, p141). In fact, with high-quality data for Qz at two or three levels, there is no reason not to use accurate integration formulae. However, practice in this area has been quite poor, as trade brochures that the author has seen use the inaccurate Mean-Section Method for integrating vertically over only three or four data points, when its errors would be rather larger than when it is used for many verticals across a channel, as described previously. This seems to be a ripe area for research.

15

River hydraulics

John Fenton

Coil for producing magnetic eld Signal probes

Figure 2-3. Electromagnetic installation, showing coil and signal probes

2.2.6 Acoustic-Doppler Current Proling methods: In these, a beam of sound of a known frequency is transmitted into the uid, often from a boat. When the sound strikes moving particles or regions of density difference moving at a certain speed, the sound is reected back and received by a sensor mounted beside the transmitter. According to the Doppler effect, the difference in frequency between the transmitted and received waves is a direct measurement of velocity. In practice there are many particles in the uid and the greater the area of ow moving at a particular velocity, the greater the number of reections with that frequency shift. Potentially this method is very accurate, as it purports to be able to obtain the velocity over quite small regions and integrate them up. However, this method does not measure in the top 15% of the depth or near the boundaries, and the assumption that it is possible to extract detailed velocity prole data from a signal seems to be optimistic. The lecturer remains unconvinced that this method is as accurate as is claimed. 2.2.7 Electromagnetic methods The motion of water owing in an open channel cuts a vertical magnetic eld which is generated using a large coil buried beneath the river bed, through which an electric current is driven. An electromotive force is induced in the water and measured by signal probes at each side of the channel. This very small voltage is directly proportional to the average velocity of ow in the cross-section. This is particularly suited to measurement of efuent, water in treatment works, and in power stations, where the channel is rectangular and made of concrete; as well as in situations where there is much weed growth, or high sediment concentrations, unstable bed conditions, backwater effects, or reverse ow. This has the advantage that it is an integrating method, however in the end recourse has to be made to empirical relationships between the measured electrical quantities and the ow. 2.2.8 Flow measuring structures These are often bound up with control and regulatory functions, as well as measurement. We will not treat them in this course.

2.3 The analysis and use of stage and discharge measurements


2.3.1 Stage discharge method Almost universally the routine measurement of the state of a river is that of the stage, the surface elevation at a gauging station, usually specied relative to an arbitrary local datum. While surface elevation is an important quantity in determining the danger of ooding, another important quantity is the actual ow rate past the gauging station. Accurate knowledge of this instantaneous discharge - and its time integral, the total volume of ow - is crucial to many hydrologic investigations and to practical operations of a river and its chief environmental and commercial resource, its water. Examples include decisions on the allocation of water resources, the design of reservoirs and their associated spillways, the calibration of models, and the interaction with other computational components of a network. The traditional way in which volume ow is inferred is for a rating curve to be derived for a particular 16

River hydraulics

John Fenton

Stage

Steady ow rating curve

Actual ood event

A measured stage value

Qfalling Qrated

Qrising

Discharge

Figure 2-4. Stage-discharge diagram showing the steady-ow rating curve and an exaggerated looped trajectory of a particular ood event

gauging station, which is a relationship between the stage measured and the actual ow passing that point. The measurement of ow is done at convenient times by traditional hydrologic means, with a current meter measuring the ow velocity at enough points over the river cross section so that the volume of ow can be obtained for that particular stage, measured at the same time. By taking such measurements for a number of different stages and corresponding discharges over a period of time, a number of points can be plotted on a stage-discharge diagram, and a curve drawn through those points, giving what is hoped to be a unique relationship between stage and ow, the rating curve, as shown in Figure 2-4. This is then used in the future so that when stage is routinely measured, it is assumed that the corresponding discharge can be obtained from that curve, such as the discharge Qcalculated shown in the gure for a particular value of stage. There are several problems associated with the use of a Rating Curve:
The assumption of a unique relationship between stage and discharge is, in general, not justied. Discharge is rarely measured during a ood, and the quality of data at the high ow end of the curve might be quite poor. It is usually some sort of line of best t through a sample made up of a number of points - sometimes extrapolated for higher stages. It has to describe a range of variation from no ow through small but typical ows to very large extreme ood events. There are a number of factors which might cause the rating curve not to give the actual discharge, some of which will vary with time. Factors affecting the rating curve include:

The channel changing as a result of modication due to dredging, bridge construction, or vegetation growth. Sediment transport - where the bed is in motion, which can have an effect over a single ood event, because the effective bed roughness can change during the event. As a ood increases, any bed forms present will tend to become larger and increase the effective roughness, so that friction is greater after the ood peak than before, so that the corresponding discharge for a given stage height will be less after the peak. This will contribute to a ood event showing a looped curve on a stage-discharge diagram as is shown on Figure 2-4. Backwater effects - changes in the conditions downstream such as the construction of a dam or ooding in the next waterway. 17

River hydraulics

John Fenton

Unsteadiness - in general the discharge will change rapidly during a ood, and the slope of the water surface will be different from that for a constant stage, depending on whether the discharge is increasing or decreasing, also contributing to a ood event appearing as a loop on a stage-discharge diagram such as Figure 2-4. Variable channel storage - where the stream overows onto ood plains during high discharges, giving rise to different slopes and to unsteadiness effects. Vegetation - changing the roughness and hence changing the stage-discharge relation. Ice - which we can ignore this is Australia, after all. Some of these can be allowed for by procedures which we will describe later.
1 High ow Low ow Flood 2 3 4 5 Distant control Channel control Larger body of water

Channel control Gauging station Local control

Figure 2-5. Section of river showing different controls at different water levels with implications for the stage discharge relationship at the gauging station shown

A typical set-up of a gauging station where the water level is regularly measured is given in Figure 2-5 which shows a longitudinal section of a stream. Downstream of the gauging station is usually some sort of xed control which may be some local topography such as a rock ledge which means that for relatively small ows there is a relationship between the head over the control and the discharge which passes. This will control the ow for small ows. For larger ows the effect of the xed control is to drown out, to become unimportant, and for some other part of the stream to control the ow, such as the larger river downstream shown as a distant control in the gure, or even, if the downstream channel length is long enough before encountering another local control, the section of channel downstream will itself become the control, where the control is due to friction in the channel, giving a relationship between the slope in the channel, the channel geometry and roughness and the ow. There may be more controls too, but however many there are, if the channel were stable, and the ow steady (i.e. not changing with time anywhere in the system) there would be a unique relationship between stage and discharge, however complicated this might be due to various controls. In practice, the natures of the controls are usually unknown. Something which the concept of a rating curve overlooks is the effect of unsteadiness, or variation with time. In a ood event the discharge will change with time as the ood wave passes, and the slope of the water surface will be different from that for a constant stage, depending on whether the discharge is increasing or decreasing. Figure 2-5 shows the increased surface slope as a ood approaches the gauging station. The effects of this are shown on Figure 2-4, in somewhat exaggerated form, where an actual ood event may not follow the rating curve but will in general follow the looped trajectory shown. As the ood increases, the surface slope in the river is greater than the slope for steady ow at the same stage, and hence, according to conventional simple hydraulic theory explained below, more water is owing down the river than the rating curve would suggest. This is shown by the discharge marked Qrising obtained from the horizontal line drawn for a particular value of stage. When the water level is falling the slope and hence the discharge inferred is less. The effects of this might be important - the peak discharge could be signicantly underestimated during 18

River hydraulics

John Fenton

highly dynamic oods, and also since the maximum discharge and maximum stage do not coincide, the arrival time of the peak discharge could be in error and may inuence ood warning predictions. Similarly water-quality constituent loads could be underestimated if the dynamic characteristics of the ood are ignored, while the use of a discharge hydrograph derived inaccurately by using a single-valued rating relationship may distort estimates for resistance coefcients during calibration of an unsteady ow model. The use of slope as well as stage Although the picture in Figure 2-5 of the factors affecting the stage and discharge at a gauging station seems complicated, the underlying processes are capable of quite simple description. In a typical stream, where all wave motion is of a relatively long time and space scale, the governing equations are the long wave equations, which are a pair of partial differential equations for the stage and the discharge at all points of the channel in terms of time and distance along the channel. One is a mass conservation equation, the other a momentum equation. Under the conditions typical of most ows and oods in natural waterways, however, the ow is sufciently slow that the equations can be simplied considerably. Most terms in the momentum equation are of a relative magnitude given by the square of the Froude number, which is U 2 /gD, where U is the uid velocity, g is the gravitational acceleration, and D is the mean depth of the waterway. In most rivers, even in ood, this is small, and the approximation may be often used. For example, a ow of 1 m s1 with a depth of 2 m has F 2 0.05. Under these circumstances, a surprisingly good approximation to the momentum equation of motion for ow in a waterway is the simple equation:
+ Sf = 0, (2.19) x where is the surface elevation, x is distance along the waterway and Sf is the friction slope. The usual practice is to use an empirical friction law for the friction slope in terms of a conveyance function K , so that we write Q2 Sf = 2 , (2.20) K in which Q is the instantaneous discharge, and where the dependence of K on stage at a section may be determined empirically, or by a standard friction law, such as the Gauckler-Manning-Strickler formula or Chzys formula:

G-M-S: K =

1 A5/3 n P 2/3

or Chzy: K = C

A3/2 , P 1/2

(2.21)

where n and C are Mannings and Chzys coefcients respectively, while A is cross-sectional area and P is wetted perimeter, which are both functions of depth and x, as the cross-section usually changes along the stream. In most hydrographic situations K would be better determined by measurements of ow and slope rather than by these formulae as they are approximate only and the roughness coefcients are usually poorly known. Even though equation (2.20) was originally intended for ow which is both steady (unchanging in time) and uniform (unchanging in space), it has been widely accepted as the governing friction equation in more generally unsteady and non-uniform ows. Hence, substituting (2.20) into (2.19) gives us an expression for the discharge, where we now show the functional dependence of each variable: q Q(t) = K((t)) S (t), (2.22) where we have introduced the symbol S = /x for the slope of the free surface, positive in the downstream direction, in the same way that we use the symbol Sf for the friction slope. This gives us an expression for the discharge at a point and how it might vary with time. Provided we know 1. the stage and the dependence of conveyance K on stage at a point from either measurement or the G-M-S or Chzys formulae, and 2. the slope of the surface, 19

River hydraulics

John Fenton

we have a formula for calculating the discharge Q which is as accurate as is reasonable to be expected in river hydraulics. Equation (2.22) shows how the discharge actually depends on both the stage and the surface slope, whereas traditional hydrography assumes that it depends on stage alone. If the slope does vary under different backwater conditions or during a ood, then a better hydrographic procedure would be to gauge the ow when it is steady, and to measure the surface slope , thereby enabling a particular value of K to be calculated for that stage. If this were done over time for a number of different stages, then a stage-conveyance relationship could be developed which should then hold whether or not the stage is varying. Subsequently, in day-to-day operations, if the stage and the surface slope were measured, then the discharge calculated from equation (2.22) should be quite accurate, within the relatively mild assumptions made so far. All of this holds whether or not the gauging station is affected by a local or channel control, and whether or not the ow is changing with time. If hydrography had followed the path described above, of routinely measuring surface slope and using a stage-conveyance relationship, the science would have been more satisfactory. Effects due to the changing of downstream controls with time, downstream tailwater conditions, and unsteadiness in oods would have been automatically incorporated, both at the time of determining the relationship and subsequently in daily operational practice. However, for the most part slope has not been measured, and hydrographic practice has been to use rating curves instead. The assumption behind the concept of a discharge-stage relationship or rating curve is that the slope at a station is constant over all ows and events, so that the discharge is a unique function of stage Qr () where we use the subscript r to indicate the rated discharge. Instead of the empirical/rational expression (2.22), traditional practice is to calculate discharge from the equation
Q(t) = Qr ((t)),

(2.23)

thereby ignoring any effects that downstream backwater and unsteadiness might have, as well as the possible changing of a downstream control with time. In comparison, equation (2.22), based on a convenient empirical approximation to the real hydraulics of the river, contains the essential nature of what is going on in the stream. It shows that, although the conveyance might be a unique function of stage which it is possible to determine by measurement, because the surface slope will in general vary throughout different ood events and downstream conditions, discharge in general does not depend on stage alone. 2.3.2 Stage-conveyance curves The above argument suggests that ideally the concept of a stage-discharge relationship be done away with, and replaced by a stage-conveyance relationship. Of course in many, even most, situations it might well be that the surface slope at a gauging station does vary but little throughout all conditions, in which case the concept of a stage-discharge relationship would be accurate. In most situations it is indeed the case that there is little deviation of results from a unique stage-discharge relationship. The use of slope in determining ow There is a considerable amount of hydraulic justication for using equation (2.22). q Q(t) = K((t)) S (t),

(2.24)

It could not be claimed that this is a theoretical justication, as they are based on empirical friction laws but, based on the cases studied above, the incorporation of slope appears to give a superior and more fundamental description of the processes at work, and handles both long-term effects due to downstream conditions changing and short-term effects due to the ow changing.

This suggests that a better way of determining streamows in general, but primarily where backwater and unsteady effects are likely to be important, is for the following procedure to be followed: 1. At a gauging station, two measuring devices for stage be installed, so as to be able to measure 20

River hydraulics

John Fenton

the slope of the water surface at the station. One of these could be at the section where detailed ow-gaugings are taken, and the other could be some distance upstream or downstream such that the stage difference between the two points is enough that the slope can be computed accurately enough. As a rough guide, this might be, say 10 cm, so that if the water slope were typically 0.001, they should be at least 100 m apart. 2. Over time, for a number of different ow conditions the discharge Q would be measured using conventional methods such as by current meter. For each gauging, both surface elevations would be recorded, one becoming the stage to be used in the subsequent relationship, the other so that p the surface slope S can be calculated. Using equation (2.22), Q = K() S , this would give the appropriate value of conveyance K for that stage, automatically corrected for effects of unsteadiness and downstream conditions. 3. From all such data pairs ( i , Ki ) for i = 1, 2, . . ., the conveyance curve (the functional dependence of K on ) would be found, possibly by piecewise-linear or by global approximation methods, in a similar way to the description of rating curves described below. Conveyance has units of discharge, and as the surface slope is unlikely to vary all that much, we note that there are certain advantages in representing rating curves on a plot using the square root of the discharge, and it my well be that the stage-conveyance curve would be displayed and approximated best using ( K, ) axes. 4. Subsequent routine measurements would obtain both stages, including the stage to be used in the stage-conveyance relationship, and hence the water surface slope, which would then be substituted into equation (2.22) to give the discharge, corrected for effects of downstream changes and unsteadiness. The effects of varying roughness Notes to be added Attempting to include unsteady effects In conventional hydrography the stage is measured repeatedly at a single gauging station so that the time derivative of stage can easily be obtained from records but the surface slope along the channel is not measured at all. The methods of this section are all aimed at obtaining the slope in terms of the stage and its time derivatives at a single gauging station. The simplest and most traditional method of calculating the effects of unsteadiness has been the Jones formula, derived by B. E. Jones in 1916 (see for example Chow 1959, Henderson 1966). The principal assumption is that to obtain the slope, the x derivative of the free surface, we can use the time derivative of stage which we can get from a stage record, by assuming that the ood wave is moving without change as a kinematic wave (Lighthill and Whitham, 1955) such that it obeys the partial differential equation:
h h +c = 0, t x

(2.25)

where B is the width of the surface and Qr is the steady rated discharge corresponding to stage , and where we have expressed this also in terms of the conveyance K , where Qr = K() S , and the slope S is the mean slope of the stream. A good approximation is c 5/3 U , where U is the mean stream velocity.

where h is the depth and c is the kinematic wave speed. Solutions of this equation are simply waves travelling at a velocity c without change. The equation will be obtained as one of a consistent series of approximations in Section 3. The kinematic wave speed c is given by the derivative of ow with respect to cross-sectional area, the Kleitz-Seddon law 1 dK p 1 dQr = c= (2.26) S, B d B d

The Jones method assumes that the surface slope S in equation can be simply related to the rate of change of stage with time, assuming that the wave moves without change. Thus, equation (2.25) gives 21

River hydraulics

John Fenton

an approximation for the surface slope: h/x 1/c h/t. We then have to use the simple geometric relation between surface gradient and depth gradient, that /x = h/x S , such that we have the approximation
h S + 1 h =S x x c t and recognising that the time derivative of stage and depth are the same, h/t = /t, equation (2.22) gives r 1 Q=K S+ (2.27) c t If we divide by the steady discharge corresponding to the rating curve we obtain r Q 1 = 1+ (Jones) Qr cS t S =

In situations where the ood wave does move as a kinematic wave, with friction and gravity in balance, this theory is accurate. In general, however, there will be a certain amount of diffusion observed, where the wave crest subsides and the effects of the wave are smeared out in time. To allow for those effects Fenton (1999) provided the theoretical derivation of two methods for calculating the discharge. The derivation of both is rather lengthy. The rst method used the full long wave equations and approximated the surface slope using a method based on a linearisation of those equations. The result was a differential equation for dQ/dt in terms of Q and stage and the derivatives of stage d/dt and d2 /dt2 , which could be calculated from the record of stage with time and the equation solved numerically. The second method was rather simpler, and was based on the next best approximation to the full equations after equation (2.25). This gives the advection-diffusion equation
h 2h h +c = 2, (2.28) t x x where the difference between this and equation (2.25) is the diffusion term on the right, where is a diffusion coefcient (with units of L2 T1 ), given by = K . 2B S

Equation (2.28), to be studied in Section 3, is a consistent low-inertia approximation to the long wave equations, where inertial terms, which are of the order of the square of the Froude number, which approximates motion in most waterways quite well. However, it is not yet suitable for the purposes of this section, for we want to express the x derivative at a point in terms of time derivatives. To do this, we use a small-diffusion approximation, we assume that the two x derivatives on the right of equation (2.28) can be replaced by the zero-diffusion or kinematic wave approximation as above, /x 1/c /t, so that the surface slope is expressed in terms of the rst two time derivatives of stage. The resulting expression is:
h 2h h +c = 2 2, t x c t and solving for the x derivative, we have the approximation S =
2 h S + 1 h d h , =S x x c t c3 dt2

22

River hydraulics

John Fenton

where Q is the discharge at the gauging station, Qr () is the rated discharge for the station as a function of stage, S is the bed slope, c is the kinematic wave speed given by equation (2.26): 1 dQr S dK = , c= B d B d in terms of the gradient of the conveyance curve or the rating curve, B is the width of the water surface, and where the coefcient is the diffusion coefcient in advection-diffusion ood routing, given by:
= K Qr = . 2B S 2B S

and substituting into equation (2.22) gives v u 1 d d2 u Q = Qr ()u |{z} + 1 3 2 u cS dt c S dt tRating curve | {z } | {z }
Jones formula Diffusion term

(2.29)

(2.30)

In equation (2.29) it is clear that the extra diffusion term is a simple correction to the Jones formula, allowing for the subsidence of the wave crest as if the ood wave were following the advection-diffusion approximation, which is a good approximation to much ood propagation. Equation (2.29) provides a means of analysing stage records and correcting for the effects of unsteadiness and variable slope. It can be used in either direction:
If a gauging exercise has been carried out while the stage has been varying (and been recorded), the value of Q obtained can be corrected for the effects of variable slope, giving the steady-state value of discharge for the stage-discharge relation, And, proceeding in the other direction, in operational practice, it can be used for the routine analysis of stage records to correct for any effects of unsteadiness.

The ideas set out here are described rather more fully in Fenton & Keller (2001). An example A numerical solution was obtained for the particular case of a fast-rising and falling ood in a stream of 10 km length, of slope 0.001, which had a trapezoidal section 10 m wide at the bottom with side slopes of 1:2, and a Mannings friction coefcient of 0.04. The downstream control was a weir. Initially the depth of ow was 2 m, while carrying a ow of 10 m3 s1 . The incoming ow upstream was linearly increased ten-fold to 100 m3 s1 over 60 mins and then reduced to the original ow over the same interval. The initial backwater curve problem was solved and then the long wave equations in the channel were solved over six hours to simulate the ood. At a station halfway along the waterway the computed stages were recorded (the data one would normally have), as well as the computed discharges so that some of the above-mentioned methods could be applied and the accuracy of this work tested. Results are shown on Figure 2-6. It can be seen that the application of the diffusion level of approximation f equation (2.29) has succeeded well in obtaining the actual peak discharge. The results are not exact however, as the derivation depends on the diffusion being sufciently small that the interchange between space and time differentiation will be accurate. In the case of a stream such as the example here, diffusion is relatively large, and our results are not exact, but they are better than the Jones method at predicting the peak ow. Nevertheless, the results from the Jones method are interesting. A widely-held opinion is that it is not accurate. Indeed, we see here that in predicting the peak ow it was not accurate in this problem. However, over almost all of the ood it was accurate, and predicted the time of the ood peak well, which is also an important result. It showed that both before and after the peak the discharge wave led the stage wave, which is of course in phase with the curve showing the ow computed from the stage graph and the rating curve. As there may be applications where it is enough to know the arrival time of the ood peak, this is a useful property of the Jones formula. Near the crest, however, the rate of rise became small and so did the Jones correction. Now, and only now, the inclusion of the extra 23

River hydraulics

John Fenton

100 90 80 70 60 Flow ( m3 s1 ) 50 40 30 20 10 0 0 1 2 3 Time (hours) Figure 2-6. Simulated ood with hydrographs computed from stage record using three levels of approximation 4 5 6 Actual ow From rating curve Jones formula Eqn (2.29)

diffusion term gave a signicant correction to the maximum ow computed, and was quite accurate in its prediction that the real ow was some 10% greater than that which would have been calculated just from the rating curve. In this fast-rising example the application of the unsteady corrections seems to have worked well and to be justied. It is no more difcult to apply the diffusion correction than the Jones correction, both being given by derivatives of the stage record. 2.3.3 Slope-Stage-Discharge Method This is essentially the method which has been proposed earlier, incorporating the effects of slope. It is presented in some books and in International and Australian Standards, however, especially in the latter, the presentation is confusing and at a low level, where no reference is made to the fact that underlying it the slope is being measured. Instead, the fall is described, which is the change in surface elevation between two surface elevation gauges and is simply the slope multiplied by the distance between them. No theoretical justication is provided and it is presented in a phenomenological sense (see, for example, Herschy 1995). An exception is Boiten (2000), however even that presentation loses sight of the pragmatic nature of determining a stage-conveyance relationship with equation (2.22), and instead uses the Gauckler-ManningStrickler formula in its classical form 1 A5/3 p Q= S , n P 2/3 where it is assumed that the discharge must be given using these precise geometrical quantities of A and P . It is rather more pragmatic to determine K() by measurements.

3. The propagation of waves in waterways


We now spend some time deriving the full equations for unsteady non-uniform ow. The fundamental assumption we make is that the ow is slowly varying along the channel. The mathematics uses a number of concepts from vector calculus, however we nd that we can obtain general equations very powerfully, and the assumptions and approximations (actually very few!) are clear. 24

River hydraulics

John Fenton

3.1 Mass conservation equation

i x

Top of control volume

Water surface z z x Q Q + Q

Figure 3-1. Elemental length of channel showing control volume extended into the air

Consider the elemental section of thickness x of non-uniform waterway shown in Figure 3-1, bounded by two vertical planes parallel to the y z plane. Consider also the control volume made up of this elemental section, but continued into the air such that the bottom and lateral boundaries are the river banks, and the upper boundary is arbitrary but never intersected by the water. The Mass Conservation equation in integral form is, written for a control volume CV bounded by a control surface CS, Z Z dV + u. dS n = 0, t CV CS | {z } | {z }
Total mass in CV Rate of ow of mass across boundary

where t is time, dV is an element of volume, u is the velocity vector, n is a unit vector with direction n normal to and directed outwards from the control surface such that u. is the component of velocity normal to the surface at any point, and dS is the elemental area of the control surface. As the density of the air is negligible compared with the water, the domain of integration in the rst integral reduces to the volume of water in the control volume, and considering the elemental slice, dV = x dA , where dA is an element of cross-sectional area, the term becomes Z A x dA = x t t
A

Now considering the second integral, on the upstream face of the control surface, u. = u , where u n is the x component of velocity, so that the contribution due to ow entering the control volume is Z u dA = Q.
A

Similarly the downstream face contribution is


Q x . + (Q + Q) = + Q + x

On the boundaries which are the banks of the stream, the velocity component normal to the boundary is very small and poorly-known. We will include it in a suitably approximate manner. We lump this contribution from groundwater, inow from rainfall, and tributaries entering the waterway, as a volume rate of q per unit length entering the stream. The rate at which mass enters the control volume is qx 25

River hydraulics

John Fenton

(i.e. an outow of qx). Combining the contributions from the rate of change of mass in the CV and the net contribution across the two faces, and dividing by x we have the unsteady mass conservation equation
A Q + = q. (3.1) t x Remarkably for hydraulics, this is an almost-exact equation - the only signicant approximation we have made is that the waterway is straight! If we want to use surface elevation as a variable in terms of surface area, it is easily shown that in an increment of time t if the surface changes by an amount , then the area changes by an amount A = B , from which we obtain A/t = B /t, and the mass conservation equation can be written Q + = q. (3.2) t x The assumption that the waterway is straight has almost universally been made. Fenton & Nalder (1995)1 have considered waterways curved in plan (i.e. most rivers!) and obtained the result (cf. equation 3.1): nm A Q + = q, 1 r t s B

where nm is the transverse offset of the centre of the river surface from the curved streamwise reference axis s, 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. It can be seen that if it is possible to choose the reference axis to coincide with the centre of the river viewed in plan, then nm = 0 and curvature has no effect on this equation. This choice of axis is not always possible, however, as the geometry of the river changes with surface height.

3.2 Momentum conservation equation


We can write the Momentum Conservation Equation in integral form, similarly to the Mass Conservation Equation: Z Z u dV + u u. dS n = P t CV CS | {z } | {z }
Total momentum in CV Rate of ow of momentum across boundary

where P is the force exerted on the uid in the control volume by both body and surface forces, the latter including shear forces and pressure forces. This is a vector equation. There are twoRmain contributions to P, forces due to pressure and shear stresses. The pressure term can be written p dS , the negative sign showing that the local force acts in the direction opposite to the n
CS

outward normal. Substituting these contributions into the momentum equation: Z Z Z u dV + u u. dS = Shear force p dS n n t
CV CS CS

The last term is very difcult to evaluate for non-prismatic waterways, as the pressure and the nonconstant unit vector have to be integrated over all the submerged faces of the control surface. A much 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
Fenton, J. D. & Nalder, G. V. (1995), Long wave equations for waterways curved in plan, in Proc. 26th Congress IAHR, London, Vol. 1, pp. 573578.
1

26

River hydraulics

John Fenton

surface integral into a simple volume integral. It can be understood by considering the uid to be divided up into a number of elemental parallelepipeds, the pressure force on one elemental face being cancelled by its force on the other. Now taking the x component of the vector momentum equation we obtain: Z Z Z p dV , u dV + u u. dS = Horizontal shear force n | {z } t x (c) CV | {z } | CV{z | } CS {z }
(a) (b) (d)

and we now make hydraulic approximations for these terms. The rst two are obtained in the same manner as for the mass conservation equation. (a) Unsteady term: The rst term is Z Z Q x. udV = u dA x = t t t
CV A

(b) Momentum ux term: There are two parts to this. The rst is from contributions on solid boundaries and the air boundary due to ow seeping in or out of the ground or from rainfall or tributaries. They are lumped together as an inow q per unit length, such that the mass rate of inow is qx, (i.e. an outow of qx) and if this inow has a streamwise velocity of uq before it mixes with the water, the contribution is
qxuq .

(3.3)

The main one is the contribution of momentum due to uid crossing the control surface: Z Z Z 2 u u. dS = n u dA + u2 dA.
U/S & D/S Face U/S Face D/S Face

(3.4)

Almost never do we know the precise velocity distribution over the face. We introduce an empirical quantity such that the effects of both non-uniformity of velocity over a section and turbulent uctuations are approximated by Z Q2 u2 dA = , (3.5) A
A

where is the momentum coefcient or Boussinesq coefcient. If the velocity were constant and steady over the section would have a value of 1, which is the usual approximation to a term which otherwise would be very difcult to evaluate. We will retain it, however, its effects are small in many situations as shown below, as it is always associated with terms which are of relative magnitude that of the square of the Froude number. Hence we write the contribution (3.4) as Z u u. dS = U 2 AU/S Face + U 2 AD/S Face n
CS

U 2 A x x Q2 x. = x A =

(c) Shear force term: The shear forces are tangential to the boundary of the stream at each point, and in the general case of a non-prismatic waterway, the geometry over which they act is complicated, not particularly well-known, and the ow structure is less well-known. Following the usual convention 27

River hydraulics

John Fenton

in river hydraulics, a convenient empirical expression is adopted instead. The approximation is made here that: Horizontal component of shear force = Weight of uid (Sf ) = g A x Sf where Sf is a small dimensionless quantity, which in derivations based on energy is the energy gradient. Here we think of it as an empirical coefcient relating the horizontal component of the friction force to the total gravitational force of the uid in the control volume, and we will call it the friction slope. The negative sign is introduced because in the usual case where ow is in the +x direction, the shear force is in the other direction. Later we will assume that it can be given by the G-M-S formula, where the local and instantaneous depth and discharge are used. (d) Pressure gradient term This is
Z

p dV. x

CV

The approximation we now make, common throughout almost all open-channel hydraulics, is the hydrostatic approximation, that pressure is given by
p g height of water above = g ( z) ,

(3.6)

which is the pressure at a point of elevation z where the free surface directly above has elevation . This is the expression obtained in hydrostatics, where the uid is not moving. It is an excellent approximation except where the ow is strongly curved, such as where there are short waves on the ow, or near a structure which disturbs the ow locally. Differentiating with respect to x, equation (3.6) gives
p = g , x x

such that the pressure gradient is determined by the downstream slope of the free surface, which we assume is constant across the stream. Substituting into the pressure term we obtain Z p dV = g A x x x
CV

Collecting terms: Combining our four contributions to the momentum equation, including the inow term from equation (3.3) and dividing by x we obtain Q2 Q + = gASf gA + quq . t x A x Expanding:
Q Q Q2 A Q + 2 2 = gASf gA + quq . t A x A x x In both mass and momentum equations we have derivatives of area. It is more convenient to use the surface elevation . We can show that for a small change in surface elevation , A = B , so that, as the bed does not move, A =B t t For the x derivative we must also allow for the fact that the bed elevation also depends on x.

The cross-section of a river in Figure ?? 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 28

River hydraulics

John Fenton

B At x At x + x z A
Z

Figure 3-2. Two channel cross-sections separated by x

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. If the change in the bed is Z as shown in the gure, R in general this varies across the section, and so the contribution to A is 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
B

In practice the precise details of the bed are rarely known, and it is convenient to introduce Z , the mean change in bed level across the section. Hence we have
A = B B Z.

Now we express this in terms of the mean bed slope across the stream S . In a distance x the mean bed level across the channel then changes by Z = S x under the water. We use the convention for bed slope that a downwards-sloping bed, the usual situation, has a positive value of bed slope. The total increase of area is A = B S x + B , giving A =B +S . x x Substituting these relations gives the momentum equation governing ows and long waves in waterways with Q and as dependent variables: Q Q Q2 B Q2 B Q + 2 + gA 2 = 2 S gASf + quq . (3.7) t A x A x A

3.3 The nature of the propagation of long waves and oods in rivers
Equations (3.2) and (3.7) are often called the Saint-Venant equations. They form a pair of partial differential equations for the surface elevation and discharge Q at any point in a stream x at any time t. These equations are used to simulate wave motions in rivers and canals, notably the propagation of ood waves and the routine simulation of irrigation channel operations. There is a software industry which specialises in numerical solutions. There have been a number of different interpretations of the nature of the propagation of long waves and oods in rivers, as described by these equations. Some quite misleading results have been obtained. We will now consider these, as they are often considered in research papers as well as more practical applications like the choice of software or, more importantly, and understanding of the real nature of long wave propagation. 3.3.1 The use of characteristics a view as waves travelling up and downstream with little

29

River hydraulics

John Fenton

diminution It can be shown mathematically that solutions of these two partial differential equations can be expressed as four ordinary differential equations. Two of the differential equations are for x as a function of t where in this case x(t) is the position of a mathematically convenient path known as a characteristic. The rst two equations are
Q dx = C, dt A where the rst part Q/A, is simply times the mean uid velocity in the waterway at that section. The second part C is the quantity r gA Q2 2 C= (3.8) + 2 , B A which is more important. The two velocities C , are relative to the water, corresponding to both upstream and downstream propagation of information in sub-critical ow. The result in equation (3.8) has included the momentum coefcient . If we assume that velocity is constant over the section such that p = 1 we obtain the traditional result that C = gA/B , where A/B is the mean depth.

It can be shown that on characteristics in given by equation (3.8), the differential equations to be satised are: d dQ Q2 B Q2 d Q Q + = 2 S gASf + q uq C . B C (3.9) A dt dt A A A dx

taking the corresponding plus or minus signs in each case. The use of characteristics has led to a widespread misconception in hydraulics and C is usually referred to as the speed of propagation of waves. It is not it is the speed of propagation of information along characteristics. Some of that informationp appear as wave motion at that speed, but not all. The will familiar and widely-quoted result C = gA/B (which from equation (3.8) implicitly contains the = 1 or = 0 approximations) has led to a common misconception in hydraulics such that C has often been referred to as the speed of propagation of waves. It is not it is the speed of characteristics. If surface elevation were constant on a characteristic there would be some justication in using the term wave speed for the quantity C , as disturbances travelling at that speed could be observed. However as equation (3.9) holds in general, neither , surface elevation, nor Q, is constant on the characteristics and one would not have observable disturbances or discharge uctuations travelling at C relative to the water. While C may be the speed of propagation of information in the waterway relative to the water, it cannot properly be termed the wave speed as it would usually be understood. 3.3.2 The low inertia approximation diffusion routing and nature of wave propagation in waterways Equations (3.2) and (3.7) (the long wave or Saint-Venant equations) are used to simulate wave motions in rivers and canals, notably the propagation of ood waves and the routine simulation of irrigation channel operations. There is a software industry which specialises in numerical solutions. We have shown that mathematically the solutions of these equations look like two families of waves, propagating at two different velocities, one upstream, the other downstream, with differing amounts of diffusive dissipation. The upstream propagating waves show rather more dissipation. The propagation in both directions is important in situations where transients are rapid, such as in hydro-electric supply canals. In most situations, however, the ow velocity is relatively small, such that a large simplication is possible. We now show that the rst two terms in equation (3.7), relative to the others, are of the order of magnitude of F 2 , and can be ignored. Firstly consider Q/t. The discharge Q is of an order of magnitude V W D, where V is a typical velocity, W is a typical width, and D is a typical mean depth. The time scale of motion is given by L/V , where L is a typical length scale of the wave motion down the waterway, and our velocity scale gives a

30

River hydraulics

John Fenton

measure of how quickly it is swept past. Hence,


V 2W D Q V W D = . is of a scale t L/V L

Now we examine the scale of the term gA/x, and here we assume that the vertical scale of our disturbances is the vertical scale of the channel D, giving
gA gW D2 D = . is of a scale g W D x L L

Now we compare the magnitudes of the two terms Q/t : gA/x and we nd that the ratio of the two terms is of a scale
L V 2W D V2 , = L gW D2 gD

the scale of the Froude number squared.

Hence, possibly to our surprise, we nd that the relative magnitude of the term Q/t is roughly F 2 , and in many ows in rivers and canals this is a small quantity and terms of this size can be ignored. Examining the second term in equation (3.7) it might be more obvious that it too is also of order F 2 . If we neglect both such terms of order F 2 , making the low-inertia approximation (and neglect the poorly-known inow term), we nd that the momentum equation (3.7) can simply be approximated by
+ Sf = 0, x

(3.10)

which expresses the fact that, even in a generally unsteady situation, the surface slope and the friction slope are the same magnitude. Now we use an empirical friction law for the friction slope Sf in terms of conveyance K , so that we write
Q2 , K2 where the dependence of K on depth at a section would be given by the G-M-S formula. Substituting this into (3.10) gives us an accurate approximation for the discharge in terms of the slope: r Q=K , (3.11) x Sf =

even in a generally unsteady ow situation, provided the Froude number is sufciently small. (Note that /x is always negative in situations where this theory applies!). This provides us with a good method of measuring the discharge - if we can calibrate a gauging station to give the conveyance as a function of surface height, then by measuring the surface slope we can get the discharge. At this point it is easier to introduce the local depth h such that if Z is the local elevation of the bottom,
= Z + h and Z h h Z = + = S0 , where S0 = , x x x x x r

in which case equation (3.11) can be written


Q=K S0 h , x

(3.12)

Now we eliminate the discharge Q from the equations by simply substituting equation (3.12) into the mass conservation equation (3.2), noting that as the bed does not move, /t = h/t, to give the single partial differential equation in the single variable h: , r ! 1 h h + (3.13) K S0 =0 t B x x The conveyance is usually expressed as a function of roughness and of the local depth h, so that we can 31

River hydraulics

John Fenton

We now have a rather simpler single equation in a single unknown. This is an advection-diffusion equation, and the nature of it is rather clearer than our original pair of equations. It has solutions which propagate at the propagation velocity shown. We write the equation as
h 2h h +c = 2, t x x

perform the differentiation in equation (3.13) and we assume that if the variation of the local depth is small compared with the overall slope so that |h/x| S0 we can write K h S0 dK h 2h + = (3.14) t x2 2B S | B{z dh} x | {z 0 }
Propagation velocity Diffusion coefcient

(3.15)

where c is a propagation speed, the kinematic wave speed, and is a diffusion coefcient (with units of L2 T1 ), which are given by S0 dK K . c= and = (3.16) B dh 2B S0 As K = 1/n A5/3 (h)/P 2/3 (h), we can differentiate to give the kinematic wave speed c for an arbitrary section. However for the purposes of this course we can consider a wide rectangular channel (h B ) such that A Bh and P B such that K = 1/n B h5/3 . Differentiating, dK/dh = 1/n 5/3 B h2/3 , and substituting into (3.16) we obtain S0 2/3 5 h . c= 3 n Now, the velocity of ow in this waterway is U = 1/n (A/P )2/3 S0 , which for our wide-channel approximation is U S0 /n h2/3 , and so we obtain the approximate relationship for the speed of propagation of disturbances in a wide channel:
5 c U, 3

(3.17)

which is a very simple expression: the speed of propagation of disturbances is approximately 1 2 times 3 the mean speed of the water. In equation (3.8) above it was stated that there is a velocity at which information travels in a waterway, r gA Q2 2 + 2 . C= B A p In all text books this is presented for = 1 such that C = gA/B , where A/B is the mean depth of the water. It is indeed the order of magnitude of the speed at which waves do move over essentially still water, but is widely used, incorrectly, to estimate the speed of disturbances in rivers and canals. It is called the dynamic wave speed, and part of waves do travel at this speed. However, provided the Froude number is small, such that F 2 1, equation (3.17) is a good approximation to the speed at which the bulk of disturbances propagate. Now to consider the effects of diffusion, if we examine the real or simulated propagation of waves in streams, the apparent motion is of waves propagating at the kinematic wave speed, but showing marked diminution in size as they propagate. We consider as a test case, a pool 7km long, bottom width of 7m, batter slopes of 1.5:1, a longitudinal slope of 0.0001, a depth of 2.1m at the downstream gate, and Mannings n = 0.02. We consider a base ow of 10 m3 s1 increased smoothly (a Gaussian function of time) by 25% up to a maximum of 12.5 m3 s1 and back down to the base ow over a period of about three hours. A computer program simulated conditions in the canal. The results presented in Figure 3-3 show that over the length of only 7km the peak discharge has decreased by about 50% and has spread out considerably in time - the wave propagation is not just a simple translation. In fact, the peak takes about 55 minutes to traverse the pool, whereas using the dynamic wave speed that time would be 30 32

River hydraulics

John Fenton

Inflow Outflow 12 Discharge (cu.m/s)

11

10 0 5 10 Time (hours) 15 20

Figure 3-3. Inow and outow hydrographs for a section of a waterway, showing effects of diffusion

minutes, while simply using the kinematic wave speed it would be 140 minutes. The wave has diffused considerably, showing that simple deductions based on a wave speed are only part of the picture. This difference might be important for ood warning operations. It is important to nd out more about the real nature of wave propagation in waterways. Here we provide a simple tool for estimating the relative importance of diffusion. If we were to scale the advectiondiffusion equation (3.15) such that it was in terms of a dimensionless variable x/L which would be of order of magnitude 1, then the ratio of the importance of the diffusion term to the advection term can be shown to be /cL. This looks like the inverse of a Reynolds number (which is correct the Reynolds number is the inverse of a dimensionless viscosity or diffusion number). Now we substitute in the approximations for a wide channel, giving: A measure of the importance of diffusion =
= K B = cL 2B S0 S0 K 0 (h) L K/K 0 (h) , and as K h5/3 this gives 2 S0 L 3 h . 10 S0 L

This is a useful result, for it shows us the effects of diffusion very simply, as h is the depth of the stream, and S0 L is the amount by which a stream drops over the reach of interest. Dropping the factor of 3/10, as our arguments are order-of-magnitude at best: A measure of the importance of diffusion Depth of stream . Drop of stream

For the example above, this is about 2.4, showing that diffusion is as important as advection, and reminding us that the problem is not the simple translation of a wave. This result is also interesting in considering different types of streams a steep shallow mountain stream will show little diffusion, whereas a deep gently sloping stream will have marked diffusion!

3.4 A new low-inertia approach Volume routing


We have introduced several approximations in deriving the advection-diffusion equation. Here we use a 33

River hydraulics

John Fenton

transformation of variables which enables us to use a single dependent variable in low-inertia routing. Derivation of equation: Consider the volume of uid upstream of a point x at a time t, denoted by V (x, t). From simple calculus, the derivative of volume with respect to distance x gives the crosssectional area: V /x = A, and as the time rate of R change of V at a point is equal to the total rate upstream at which the volume is increasing, which is x q dx less Q, the volume rate which is passing R the point, we have V /t = x q dx Q. Substituting the relations Z A = V /x and Q = q dx V /t (3.18)
x

into the mass conservation equation (3.1):


A Q + =q t x

shows that it is identically satised! This might have been expected, as the equation is a mass conservation equation, and hence for an incompressible uid it is a volume conservation equation. We have been able to express both A and Q in terms of a single variable V . Now we go on to use this in the simplied momentum equation (3.10):
+ Sf = 0. x Firstly, the derivative of the cross-sectional area can be related to the derivative of the stage by A =B +S , x x

so that the simplied momentum equation becomes


1 A = S Sf , B x

(3.19)

such that if we use the frictional law in general form Sf = Q2 /K 2 :


1 A Q2 + 2 S = 0, B x K

(3.20)

and substituting for Q and A in terms of V , from equation (3.18) and as both breadth B and conveyance K can be written as functions of area we obtain the single equation in the single variable s Z V 1 2V + K(Vx ) S = q dx, (3.21) t B(Vx ) x2 x in which the only approximation relative to the long wave equations has been that we ignore terms of O(F 2 ), such that it will be accurate for F 2 1. This equation, which we term the Volume Routing Equation, might be useful in a range of hydrologic and hydraulic computations, replacing the solution of the long wave equations. It is a nonlinear partial differential equation which is a single equation in a single variable. From it, deductions can be made about the nature of wave propagation in waterways, which are not as misleading as those from the characteristic formulation of the long wave equations. Relation to conventional advection-diffusion equations: In the formulation of (3.21) it is not obvious that the volume routing equation is of an advection-diffusion nature. We show that here. Consider small perturbations about a uniform ow of area A0 and discharge Q0 :
V = A0 x Q0 t + v(x, t),

34

River hydraulics

John Fenton

where is a small quantity. Immediately we have


V v = Q0 + , t t V v = A0 + , x x

and

2V 2v = 2, x2 x

giving

from a Taylor series expansion about the uniform ow, which has introduced this advective term. We do the same for B(Vx ) but recognise that we only need it to lowest order (it multiplies a Vxx term):
B(Vx ) = B (A0 ) + . . . ,

v dK v + . . . , = K (A0 ) + K(Vx ) = K A0 + x dA 0 x

and now we consider the term s s 2V 1 2V 1 = S S 1 B(Vx ) x2 SB(Vx ) x2 s p 1 2v 2v 1 1 1 1 = 2 = S 2 + ... S x 2 SB(A0 ) x SB(A0 ) from the binomial theorem. Substituting all these linearising approximations into (3.21) with q = 0, p 1 v dK 2v v + . . . S 1 1 + K (A0 ) + Q0 + 2 + . . . = 0, t dA 0 x 2 SB(A0 ) x

and multiplying out and dropping all terms in 2 and higher, we get p p 1 2v v p dK v + SK (A0 ) 1 + S 2 = 0. Q0 + K (A0 ) S + t dA 0 x 2 SB(A0 ) x Now, however, as Q0 = K (A0 ) S , the rst terms cancel, and we are left with all terms of the order of magnitude , which we divide through by to give p 2 K v dK v = (A0 ) v = 0, + S t dA x 2 SB(A0 ) x2 | {z 0 } | {z }
c0 0

which is precisely the advection-diffusion equation that we obtained previously when we assumed a base uniform ow from the outset. Initial conditions: In general for such problems as we might want to use this for, such as real rivers, the initial conditions are important. If we consider the steady state form of equation (3.21) we obtain the condition (for no inow) (note that V /t is not zero for steady ow, as volume is continually passing, but V /t = Q0 where Q0 is the initial steady ow. Then we have (more easily from equation 3.20): 2 d2 V Q0 . = B(Vx ) S (3.22) dx2 K 2 (Vx ) which is a second-order ordinary differential equation which we have to solve numerically. Note that this is just a low-inertia version of the gradually-varied ow equation
S Sf dh = . dx 1 F 2

This is a bit messier than the conventional formulation, as we have to solve the second-order equation numerically, which is usually done by introducing a subsidiary variable dV /dx which is A in our formulation! 35

River hydraulics

John Fenton

Boundary conditions: At an upstream boundary x0 we might have a given inow as a function of time Q(x0 , t), which, from equation (3.18) gives us what dV /dt is there, such that we have also to solve the ordinary differential equation dV (x0 , t)/dt = Q(x0 , t) there as part of the solution, which is relatively simple. At control points, such as a downstream structure we will usually have some relationship between Q and A such as provided by weir formulae, which gives V /t as a function of V /x there. As part of the solution we will have to differentiate numerically to give the latter and then integrate to give the updated value of V . At open boundaries, where there is no control point, we may simply be able to apply the partial differential equation as if it were an interior point, although if nite differences were being used to evaluate the spatial derivatives a different formula in terms of points to one side would have to be used. Some results and insights into the real and simulated nature of wave motion: Consider as a test case a pool 7km long, bottom width of 7m, batter slopes of 1.5:1, a longitudinal slope of 0.0001, a target depth of 2.1m at the check gate, and Mannings n = 0.02. To perform the simulation we adopt the general conditions of their Test 2-1, with an initial ow of 10 m3 s1 . The inow was increased by 25% in 15 minutes. We developed a program to solve the volume routing equation using similar approximation methods to the full model, with cubic spline approximation along the canal and simple Euler forward time stepping but with Richardson extrapolation. We embedded it in a full model of the long wave equations. The two programs simulated conditions in the canal for several hours, with an overshot weir at the check.

12 Discharge (cu.m/s)

11

Inflow Dynamic model Volume routing 10 0 0.5 1 1.5 Time (hours) 2 2.5 3

Figure 3-4. Discharge hydrographs showing inow and the outow calculated from the full equations (3.2) and (3.7) and from the volume ow routing equation (3.21).
8 7 Relative water level (cm) 6 5 4 3 2 1 0 0 0.5 1 1.5 Time (hours) 2 2.5 3 Dynamic model Volume routing

Figure 3-5. Variation of water level at the gate calculated from the full equations (3.2) and (3.7) and from the volume ow routing equation (3.21).

To demonstrate the behaviour of the canal initially the ow from the headworks was increased uniformly 36

River hydraulics

John Fenton

over 15 minutes and a constant downstream gate opening was maintained. After the ow and surface level started to increase downstream, the check gate was brought up to the required full ow in 15 mins under idealised control, and thereafter required delivering precisely the required increased ow. Figure 3-4 shows the resulting ow hydrographs. It can be seen that in this canal with a relatively mild slope and moderate friction that the outow hydrograph is very different from the inow hydrograph. The effects of the diffusion-like term with the second derivative in the volume routing equation (3.21) are strong. The time when the dynamic model rst showed some effect downstream (about 1/2 hour) corresponded closely to the calculated travel time of a dynamic wave, showing that the forerunner of the motion was a dynamic wave. However, the bulk of the motion is a relatively slow-moving kinematic-diffusion wave, which the approximate model closely predicts. Until the downstream gate was opened at 1.5 hours, only about half of the increased ow had arrived. Calculations based on the kinematic wave speed showed an expected travel time of about 2.25 hours, and from the gure it is clear that this is a more representative travel time for the whole increase of ow. What is also obvious, of course, is how efcacious the opening of the downstream gate was in bringing the ow up to required levels. Relying on the slow movement of the ow transients is not enough. In our simulation, as described, we suddenly opened the gate and thereafter maintained the desired ow. Under such conditions, what might now be a concern is the behaviour of the water level at the gate. This is shown in Figure 3-5. It can be seen that there is an initial period, corresponding to a time of rapid changes in the ow, when there was a noticeable disagreement between the two models. However, the approximate model did describe well the main feature of the ow, the increase of ow and surface height as the slow-moving kinematic wave approached. After the gate opened suddenly, as one would expect, the level was quickly drawn down, and again this is described by the approximate model. The volume routing method seems to be capable of simulating the behaviour of the pool, with some errors where rapid ow changes occur, but the overall behaviour of the movement of water masses and surface level behaviour are described satisfactorily. Numerical simulation using the volume routing equation (3.21) is very much simpler than that using the full equations.

4. Computational hydraulics
4.1 The advection equation
We will consider some special cases of the advection-diffusion equation, as they provide us with a number of insights. Consider the equation with no diffusion, known as the advection equation.:
+ u (x, t) = 0, t x

(4.1)

where (x, t) is some passive scalar, and u (x, t) is a velocity, possibly a wave speed. A typical problem is to solve the advection equation when we know (x, 0), that is, the distribution of at some initial time, and we also know what (0, t) is, namely how it is varying at a boundary. We want to obtain the solution for all x and t. 4.1.1 Exact solution for constant velocity In the case of a constant velocity u(x, t) = U , the equation has a simple analytical solution (x, t) = f (x U t), where the function f (t) is given by the history of at the upstream boundary, f (t) = (0, t), and to obtain the value at any general place and time (x, t) we just substitute f (x U t). The solution corresponds to a simple wave travelling at a speed of U . We can easily verify that this is the solution, for
t x = = df (x U t) = U f 0 (x U t) d(x U t) t df (x U t) = f 0 (x U t), d(x U t) x

37

River hydraulics

John Fenton

where f 0 (x U t) = df (x U t)/d(x U t). Substituting these values into equation (4.1) shows that it is satised exactly. Figure 4-1 shows the exact solution of a triangular wave being advected with no diffusion.
1.2 1 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 x 0.8 1 1.2

Figure 4-1. Exact solution of advection equation for triangular wave

4.1.2 An advective numerical scheme In situations where the velocity is not constant, then numerical solutions have to be made. It is rare that such a simple equation has to be solved numerically, but here we include numerical schemes as models for rather more complicated problems. The previous exact solution scheme suggests the following scheme: where the O(2 ) means that neglected terms are of the order of 2 . This is an advective scheme, which attempts to build in the nature of the solution. It can be interpreted as To obtain the solution at some point x at a later time t + , take the known value of the velocity at (x, t), namely u(x, t), and at a distance upstream given by this velocity times the time step, interpolate the value. In the case of a constant velocity u(x, t) = U this would be exact, for the value at (x, t + ) is precisely that which was upstream at (x U , t). However, if the velocity is variable, it is not exact, and errors are proportional to the square of the time step. Such advective schemes are to much to be preferred in uid mechanics, hydraulics, and hydrology. Schemes which do not incorporate the advective nature of the solution can have some very unpleasant characteristics, as we now demonstrate. 4.1.3 The simplest and most obvious nite difference scheme: Forward Time, Centre Space (FTCS) Finite difference approximations to derivatives are used throughout engineering to provide numerical solutions of partial differential equations. Here, instead of using the advective scheme above we adopt the typical types of approximations to the derivatives used in nite difference approximation. Using the forward time and centre space approximations, we substitute into the advection equation (4.1):
(x, t + ) (x, t) (x + , t) (x , t) + u (x, t) = 0, 2 {z } {z } | |
Centre approximation to /x

(x, t + ) = (x u(x, t), t) + O(2 ),

Forward approximation to /t

38

River hydraulics

John Fenton

and rearranging gives the FTCS scheme for computing the updated value at (x, t + ):
(x, t + ) = (x, t) u (x, t) [ (x + , t) (x , t)] , 2

(4.2)

so that the scheme can be represented as calculate the centre difference approximation /x ( (x + , t) (x , t))/2 , and then calculate the change in value at x by calculating the distance u and the change u /x. This can be interpreted as in Figure 4-2.
u
Updated value

Approximation to derivative at x

x+

Figure 4-2. Interpretation of FTCS scheme showing how (x, t + ) will be greater than (x, t)

We have deliberately drawn this such that the quantity has a maximum at x. This more clearly shows that when the solution is updated, the value at t + will be greater than the previous maximum. This suggests that the scheme will be unstable, as maxima will grow. This phenomenon is well-known in numerical methods for solving partial differential equations. Figure 4-3 shows such a numerical solution for an initially triangular distribution for C = u/ = 0.75, the same problem as in Figure 4-1, but here solved numerically. The parameter C is an important one in computational hydraulics, the Courant Number, which expresses how far the solution should be advected in a single time step relative to the space step. In this case, the solution should be carried 3/4 of a space step in a time step. We have found that this simple and obvious scheme is unstable, and is unable to be used at all, as was suggested by Figure 4-2.
2.0

1.5 1.0

0.5

0.0 0.0 -0.5 -1.0 0.2 0.4 0.6 0.8 1.0 1.2

-1.5 x

Figure 4-3. Unstable numerical solution with FTCS scheme and C = 0.75

4.1.4 Summary of schemes for the advection equation

39

River hydraulics

John Fenton

Figure 4-4. Physical representation of three computational schemes for solving the advection equation.

The advection equation as presented in equation (4.1) is


+ u (x, t) = 0, t x

(A1)

and we have shown that solutions to this show the behaviour of a travelling wave. Scheme 1 below approximates this behaviour. We will see, however, that most schemes (such as 3 and 4 below) do not build this behaviour in, but rather just approximate the time and space derivatives. In Figure 4-4 are shown the physical natures of common approximation schemes for the advection equation. At a certain time step t three computational values at x , x, and x + are shown by solid circles. At the next computational point time t + three updated values of the solution at x are shown, according to three schemes described below, each taking information from x u. 1. Advection scheme: The advection scheme
(x, t + ) = (x u(x, t), t),

(A2)

incorporates the travelling wave behaviour, saying interpolate the solution at a point x u, and that will be the updated value of at x at time t + . An interpolation of the points is shown by the solid curve. The value of at (x u, t) is obtained from the interpolation, which, from equation (A2), gives the value at (x, t + ).This updated solution is shown in Figure 4-4 by an open circle with a solid line. It can be shown that the scheme is always stable as the scheme mimics the real physical behaviour expected, of a translating wave. The interpolation can be done by any scheme a simple one here would be to t a quadratic to the three solid points shown. The lecturer prefers using cubic splines, which are a very powerful way of using a series of cubics to do the approximating. 2. Forward time-stepping (FT) schemes: Most other schemes do not exploit the travelling-wave nature of the solutions, but rather just approximate all the derivatives of the partial differential equation. Forward time stepping schemes all approximate the time derivative as shown in equation (4.2):
(x, t + ) (x, t) + u (x, t) 0, x

but they vary as to how the space derivative is evaluated. This can be re-written as the scheme
(x, t + ) (x, t) u (x, t). x

(A3)

This scheme can be interpreted as the change in is equal to u times the approximation to the derivative, or, travel along the line with gradient that of the approximation back a distance u, and that is the updated value. It is interesting that this is the rst-order Taylor expansion in x of the potentially-exact scheme, equation (A2). Now we consider two such schemes. These are traditionally much more common throughout computational hydraulics than advection schemes. One wonders why. 3. FTCS and other schemes which approximate the derivative accurately: Here we consider a family of schemes which approximate the derivative in x accurately, including the Forwards-TimeCentred-Space scheme considered in 4.1.3. While that spatial approximation is accurate, the time40

River hydraulics

John Fenton

stepping scheme is only rst-order, and not particularly accurate. To evaluate the derivative accurately a high-order scheme using splines or Fourier series or a centred space scheme could be used as in 4.1.3. For our purposes here it doesnt matter which scheme is used. In Figure 4-4 at x the long-dashed line shows such an accurate approximation to the local gradient of the curve, and the long-dashed circle shows the solution taken from such a scheme. It can be seen that the updated solution, near the function maximum that we are examining, is actually higher, and the wave has grown in magnitude. This suggests that such a scheme is unstable which we can show mathematically is the case. In fact, all such schemes are unuseable for any value of u because in Figure 4-4 it can be seen that the tangent is always above the interpolating function! 4. FTBS scheme: Finally we consider a simple Forwards Time Backwards Space scheme, where the derivative is approximated by a backwards difference approximation, as shown in Figure 4-4. The short-dashed line shows the backwards difference approximation to the gradient, and the corresponding updated point is the short-dashed circle. It can be seen that the solution is now lower than the accurate advection solution. This shows the phenomenon of numerical diffusion, due to such a poor level of approximation. To a lesser degree it occurs in many computational schemes. If we were free to choose u = the solution would be exact, as the point we would update from is the exact solution at x . However, u is usually a function of time and space and this cannot be satised at all points. If we were to take u > , then as can be seen on Figure 4-4 the gradient line is now above the exact solution, and the scheme would be unstable. It is common for this limitation to occur in computational schemes. It is called the Courant-FriedrichsLewy criterion, and introduces the Courant number C :
C= u 6 1 for stability,

whose essential meaning is for stability, the computational wave in a single time step should not travel more than a single space step.

4.2 The diffusion equation


The diffusion equation, obtained when the advective velocity of the medium is zero, is
2 = 2, t x

(Diffusion Equation)

and is well-known to describe many physical quantities in nature, including the ow of heat. The commonly-used Forward Time Centre Space scheme: The best-known numerical scheme is where the time derivative in the diffusion equation is approximated by a forward difference, and the diffusive term by a centre-difference expression. We obtain
(x, t + ) (x, t) (x + , t) 2 (x, t) + (x , t) = , 2

which gives the scheme


(x, t + ) = D (x , t) + (1 2D) (x, t) + D (x + , t) ,

(4.3)

in which D is the computational diffusion number D = / 2 . This is widely used, notably in civil engineering, to solve the consolidation equation in Geomechanics, which is simply the diffusion equation. It is shown in the Appendix that this has a conditional stability, such that
D= 1 6 2 2

for stability. In fact, in most cases this criterion is not that which governs the time step, rather, it is accuracy considerations. Insight into this is gained by considering Figure 4-5 which shows the results 41

River hydraulics

John Fenton

for an initial concentration of 1 at the centre point and 0 for all others (a nite equivalent to Figure A-4). Taking a value of D = 1/8 and performing four steps (to t/ 2 = 1/2) of the scheme (4.3) gives the diffused but peaked shape shown, which we have seen above is in accordance with what we expect. In one step of the limiting case D = 1/2 to the same point in time the solution has snapped through far too much and gives a physically nonsensical result shown. Clearly, accuracy rather than stability determines the desirable step size. Of course, there are many other schemes which could be tried.

Figure 4-5. Results for solving the diffusion equation numerically

4.3 Advection-diffusion combined


We consider some of the fundamentals and some possible methods. 4.3.1 Forward Time Centred Space scheme For the full advection-diffusion equation we obtain
(x, t + ) (x, t) (x + , t) (x , t) (x + , t) 2 (x, t) + (x , t) + u(x, t) = , 2 2

which gives the scheme 1 1 (x, t + ) = D + C (x , t) + (1 2D) (x, t) + D C (x + , t) , 2 2 in which C is the Courant number C = u/ , expressing the relative amount by which the solution is advected in a time step, and D is the computational diffusion number D = / 2 , expressing the effect of diffusion. Performing a Von Neumann stability analysis (such as outlined in the Appendix), after considerable difculty, it can be shown that for stability, two criteria are obtained. The rst is a limitation on the computational number D: 1 2 6 2, which is independent of the ow velocity, and is the same as we obtained above for pure diffusion as well. The second becomes u2 6 2, and it can be seen how difcult and strange the behaviour of advection and diffusion can make numerical schemes. To satisfy the rst criterion, the time step allowed is inversely proportional to diffusion, the more diffusion, the smaller the time step, which feels reasonable. However, to satisfy the second criterion, the allowable time step is proportional to the amount of diffusion, thus, strangely, the less diffusion 42

River hydraulics

John Fenton

there is, the smaller is the time step allowed for stability, and in the limit of vanishing diffusion, the scheme is unconditionally unstable, as we have already discovered! 4.3.2 A simple advection-oriented scheme The previous results suggested that the combination of advection and diffusion can be difcult to compute. Many of these difculties are overcome if the advective nature of solutions are incorporated. A simple scheme which the lecturer advocates is simply using the advective nature, writing the scheme 2 (x, t + ) = 1 + 2 (x u, t) x 2 = (x u, t) + 2 (x u, t) . x This can be interpreted as interpolate to nd the value of upstream a distance u as well as its second derivative there, and combine them as shown to give the updated value. Stability analysis: Using the von Neumann stability analysis (see the Appendix), we obtain a stability criterion which is similar to that for the pure diffusion equation. By incorporating advection exactly we have overcome any difculties with the combination of advection and diffusion as we found above. If we used the FTCS scheme for the diffusion part we would obtain the same criterion as for the pure diffusion case. 4.3.3 An advection-oriented scheme the Holly-Preissmann method This is described in some of the references, such as Sauvaget (1985), but is rather too complicated to present here.

5. Water quality
5.1 Useful sources for further reading
Table 1.4 shows some of the many references available, some which the lecturer has referred to in these notes or in his work. Those at the head of each section tend to be more important, accessible etc., and for books the University of Melbourne Library and Reference Numbers are given. The details are given in the References section at the end of these notes.

5.2 Water quality characteristics


There are many characteristics which can be used to describe the quality of water in a river: Physical: Inorganic: Organic: Gases: Biological: temperature, total dissolved solids (TDS), suspended solids, turbidity, colour and odour. pH, hardness, conductivity, nutrients, heavy metals and trace elements. biochemical oxygen demand (5 day and ultimate BODu ), chemical oxygen demand. oxygen O2 , carbon dioxide CO2 , hydrogen sulphide H2 S. micro-organisms including Escherichia coli (E-coli), pathogenic organisms and algae.

This list is not exhaustive. Acceptable standards for water quality vary depending on whether water is to be used for domestic, agricultural or industrial purposes; or whether the aim is to maintain the biology of a river.

5.3 Types of pollutant


The following is substantially taken from Fischer et al. (1979):

43

River hydraulics

John Fenton

The following list is arranged in order of hazard, starting with the least dangerous, however for Australian conditions, what is least hazardous might be among the most environmentally damaging, namely common salt, and we have modied some comments. Natural inorganic salts and sediments: These materials are not toxic in small concentrations and only become possible pollutants in excessive doses. In the case of common salt in Australia, this is a very important case. Waste heat: Once-through cooling systems for electricity generating plants use water for carrying away large quantities of low-grade waste heat. Aquatic life can be severely affected for example the extensive investigations for minimising effects of waste heat from Huntly Power Station on the Waikato River in New Zealand. Organic wastes: Domestic sewage containing ecosystem materials such as carbon, nitrogen, and phosphorous, can cause bad smells and nuisances. After treatment and dispersion (of varying degrees) it is considered acceptable to assimilate these materials into large water bodies. The biochemical oxygen demand (BOD) may be sufciently reduced so that it can be satised by the natural dissolved oxygen in the water body. Trace metals: These, often heavy metals such as lead, mercury, and cadmium, are naturally present in small amounts, but wastewater can have high concentrations. Synthetic organic chemicals: These are slow to degrade in the environment and are often bioaccumulated in the food chain. Wastewaters may have high initial dilution, but the food chain can multiply the concentration by several orders of magnitude. Biological processes can do the opposite of the physical process of turbulent mixing which reduces the concentration. Radioactive materials: Long term storage of these is necessary, and the possibility of leaking into water must be guarded against severely. In the northern hemisphere, salt mines are often chosen. The northern hemisphere also likes the look of dry parts of the southern hemisphere. Chemical and biological warfare agents These are designed to be toxic at very small doses. They have a habit of returning to damage the agents who used them (Agent Orange).

5.4 Mass balance concepts


Regulatory systems usually distinguish between point and non-point sources of pollutants. A point source is usually the discharge from a structure which is designed for the outow of wastewater. Exceptions include the accidental spill of oil from a ship and the release of radioactive wastes from a power plant. Most laws and regulations for water pollution control concern point sources. It is not usually possible to treat pollution from non-point sources. Examples include the runoff of salts and nutrients from agriculture, soil erosion, acid rain, and street drainage.

5.5 Impacts of human works


A partial list of damage caused by traditional approaches is:
Human-made reservoirs may cause deterioration in water quality because of summertime thermal stratication associated with oxygen depletion in the lower layers. The reservoirs raise the head on groundwater which may cause it to seep to the surface and bring pollutants. Diversion of water for various uses or to other watersheds reduces river ow and its ability to provide ushing and to provide a satisfactory environment for organisms.

44

River hydraulics

John Fenton

Conveyances such as canals can transport large quantities of dissolved salts, sediment, nutrients and parasites to places that would not otherwise receive such doses. Agricultural drainage systems may greatly accelerate the leaching of nutrients and salts from the land into natural hydrologic systems. Breakwaters for harbours interfere with natural circulation which could otherwise carry away pollutants. Estuarine modications can radically change circulation patterns with dire consequences for ushing of pollutants Most coastal protection works of a hard nature cause environmental damage somewhere, even if they perform the local task they were required. Sea-walls and groynes often lead to the loss of sand from beaches, an environmental problem of another kind.

Sometimes it can be very difcult to turn around traditional culture, especially where that other ubiquitous substance money is concerned. However tradition and personal interest are powerful factors as well. Morgan (1971, Monash University: Hargrave-Andrew Library 624 M847D) describes the entrenched culture of the US Army Corps of Engineers, responsible for many works in the US. Goldsmith and Hildyard (1992, UniMelb Engin f 333.731 GOLD) describe at length the effects of dams around the world.

5.6 Transport processes


Advection: Transport by an imposed current system, as in a river or coastal waters.

Convection: Vertical transport induced by hydrostatic instability, such as the vertical ow above a plain on a hot day. Diffusion (molecular): The scattering of particles by random molecular motions, which may be described by Ficks law and the classical diffusion equation. Viscosity is an effect of molecular diffusion. Diffusion (turbulent): The random scattering of particles by turbulent motion, considered roughly analogous to molecular diffusion, but with eddy diffusion coefcients, much larger than molecular diffusion coefcients. Shear: The movement of uid at different velocities at different positions.

Dispersion: The scattering of particles or a cloud of contaminant by the combined effects of shear and transverse diffusion (to be explained later). Mixing: Diffusion or dispersion as described above; turbulent diffusion in buoyant jets and plumes; any process which causes one parcel of water to mingle with or be diluted by another. Evaporation: Radiation: The transport of water vapour from a water or soil surface to the atmosphere. The ux of radiant energy, such as at a water surface.

Particle settling: The sinking (or rising) of particles having densities different from the ambient uid, such as sand grains or dead plankton. Particle entrainment: The picking up of particles such as sand or organic detritus from the bed of a water body by turbulent ow past the bed.

45

River hydraulics

John Fenton

For pollutant analysis, uctuations and irregularities are just as important as the mean ows.

5.7 Tools for problem solving


5.7.1 Order of magnitude analysis Fischer et al. (1979, p12) write: For any mixing problem a skilful analyst should be able to work out a rough approximation for the solution within a fraction of an hour!. Such a process is known as order of magnitude analysis. It is reminiscent of the wager that the physicist Richard Feynman (Feynman 1985) used to make, that he could solve any physical problem to within 10% in one minute! Such problemsolving skills are important for environmental engineers. Problems can be broken up into sub-models. Much time and effort which might otherwise be wasted can be saved by this process. 5.7.2 Numerical techniques A numerical calculation can be no better than the validity of the underlying approximations made when representing a complex process by mathematical equations. It is still necessary to make assumptions based on judgement and insight, and to leave out those processes which have little effect on the results. A wide variety of numerical methods can be applied to mixing problems, which usually means the numerical solution of differential equations, whether ordinary or partial. Typically, nite difference or nite element methods are used, although the fundamental equation for transport by a ow where diffusion occurs, the advection-diffusion equation, shows a number of surprises and pathologies when numerical solution is attempted. Stochastic (Monte Carlo methods) may be used to simulate diffusion, which is the like simulating the like. In general, the methods which are most robust and successful are those where the solution method incorporates or mimics the nature of solutions. 5.7.3 Hydraulic models These are necessary in some situations where the physical problem is sufciently complicated. Also, they reveal physical phenomena which the relatively simple equations used cannot reveal. Various phenomena revealed by experiments include: large scale vortices, internal waves and hydraulic jumps (caused by uid stratication), multilayer shear ows, blocking, doubly-diffusive convection, etc.. 5.7.4 Field studies In cases of dispersion of pollutants it is often difcult to model all the physical scales adequately. Field experiments are very important. Sometimes Lagrangian-type experiments, such as following drogues or drifters (or in the case of a famous experiment, two oating pieces of parsnip) to track ow trajectories and dispersion. However, tracking is difcult, and xed-location apparatus gives data at a reasonable cost. 5.7.5 Mixed approaches For large complicated situations a careful interweaving of all the above is probably the best way to proceed. Each piece of a problem should be done in the most practical way bearing in mind the aphorism that there is nothing quite so practical as a good theory. The nal synthesis of a varied approach will undoubtedly be better than by depending on only one approach. Each different approach illuminates the others.

5.8 A simple river model organic wastes and self purication


5.8.1 Dissolved oxygen and Biochemical Oxygen Demand (BOD) The presence of dissolved oxygen in river water is essential for maintaining plant and animal life in a river. When wastes (such as domestic sewage) containing organic material enter a river, they are 46

River hydraulics

John Fenton

decomposed (oxidised) by millions of bacteria. These depend for their survival on a good supply of oxygen in the water. This natural purication process reduces the dissolved oxygen level in the river until the rate of oxygen absorption from the air (and from oxygen-producing aquatic plants) exceeds the rate of oxygen demand from the waste. Dissolved Oxygen (O2 ): The concentration of dissolved oxygen is a measure of the health of a river and low oxygen concentration is the most important indicator of the presence of easily decomposed organic waste material. Continuous recording of the O2 content can yield as much signicant information as a stage recording gauge. Saturation concentration, Cs , of dissolved oxygen depends on temperature, and for fresh water, Elmore & Hayes (1960) developed the formula
Cs = 14.652 0.41022 T + 0.0079910 T 2 0.000077774 T 3

(5.1)

where Cs is given in mg l1 or g m3 and T is the stream temperature in o C. Some typical values given by this formula are shown in Table 5-1. Values of Cs decrease with increasing salinity (total dissolved solids) and for seawater, values are approximately 20% less than the values given here.
Table 5-1. Dependence of saturation concentration on temperature T (o C) Cs 5 12.79 10 11.27 15 10.03 20 9.02 25 8.18

Since oxygen is essential to aquatic life it is important to set standards for the minimum level of dissolved oxygen necessary to maintain the biology a particular river reach. A value of 5 g m3 might be typical, although for colder rivers, such as trout streams, the standard level could be increased to 6 g m3 . Biochemical Oxygen Demand (BOD): The impact of a pollutant on a river is normally measured as an oxygen demand (usually BOD) which is a gross measure of the concentration of oxidisable organic material. Most organic materials are biodegradable and the amount of oxygen used in the metabolism of biodegradable organics is known as the biochemical oxygen demand or BOD. The units of BOD are mg l1 or g m3 . There are two phases of the BOD reaction. The rst involves the oxidation of carbonaceous organic material, with the end products being carbon dioxide (CO2 ), ammonia (NH3 ) and water (H2 O). The second phase includes the biological oxidation of ammonia to nitrate. This second (or nitrication) phase begins much later than the rst phase and the oxygen demand is lower than the normal process of re-aeration in nature. These are shown in Figure 5-1. Unless specied otherwise, BOD values denote a measure of oxygen required for oxidation of carbonaceous organic material only. Traditionally the standard BOD test has been the 5 day BOD, (BOD5 ). This test gives a measure of the amount of oxygen required over ve days by bacteria involved in oxidising a sample of a particular waste under laboratory conditions at 20o C. Although the 5 day, 20o C BOD (BOD5 ) is widely used, it has a number of serious deciencies. The 5 day period does not usually correspond to the point where all of the waste has been consumed. It was adopted at the beginning of the century in the United Kingdom where rivers do not have a ow time to the sea greater than ve days, and summer temperatures are less than 20o C. In practice the ultimate BOD, BODu , is of most interest. Sometimes BOD20 is used to estimate BODu . 5.8.2 Biological self purication the oxygen sag model of Streeter & Phelps Biological self-purication implies that a body of water such as a river is automatically able to absorb and remove the organic compounds of efuents or wastes. Here the properties of the water revert to the conditions upstream of the waste input point after a certain length of ow or travel time. The methods of analysis used to determine the capacity of a river to assimilate organic pollution essentially began 47

River hydraulics

John Fenton

Figure 5-1. Typical oxygen demand curve of aerobic decomposition of organic matter (after McGauhey, 1968)

with the simple classical model of Streeter and Phelps in 1925. In the Streeter-Phelps model, the BOD and dissolved oxygen proles along a river are based on the assumption that there are only two major processes taking place:
bacterial oxidation of the organic matter, which is responsible for the removal of BOD and oxygen; and reaeration at the water surface.

Firstly we consider the processes at work. Deoxygenation in Rivers: The oxygen in rivers is depleted by the bacterial oxidation of suspended and dissolved organic matter discharged into them. (There can also be oxygen depletion by the organisms found in the benthic - bottom - deposits on a stream bed, but this aspect will be neglected here). The amount of oxygen required to stabilise a waste is normally measured by a BOD test, and BOD is therefore the primary source of oxygen depletion in a river. Here we assume a simple rst order model, that the rate of uptake of oxygen (rate of deoxygenation) is proportional to the amount of organic matter available:
d (BODr ) = k1 (BODr ) dt

(5.2)

where BODr g m3 is the BOD remaining at time t, a measure of the content of easily degradable organic substances, and k1 (d1 ) is the coefcient giving the rate of BOD deoxygenation.

Typical values of k1 are 0.2 to 0.35 d1 . It depends on the state of the waste material and the degree of treatment. For example, for raw municipal waste water k1 might typically be 0.35, while for treated waste water k1 might equal 0.25. The values of k1 are also temperature dependent. The temperature effect on k is usually expressed as
k1 (T ) = k1 (20) T 20

where T is the temperature and k1 (20) is the value at 20o C. The values of range from 1.03 to 1.05 in the range of 15o to 35o C. Solving the differential equation (5.2) by separation of variables, we obtain
BODr = A ek1 t ,

48

River hydraulics

John Fenton

where A is a constant. We obtain this from the initial condition such that when t = 0, BODr = BODu , the initial value, giving
BODr = BODu ek1 t .

(5.3)

In the standard laboratory test, values of the oxygen used at the end of specied intervals of time are determined. Using equation (5.3), and noting that oxygen used is given by BODu BODr , the values of k1 and BODu can be determined from laboratory data. We introduce the symbol rD for the rate of deoxygenation, which will be a negative quantity. From equations (5.2) and (5.3):
rD = d (BODr ) = k1 (BODr ) = k1 BODu ek1 t dt

(5.4)

Reoxygenation in Rivers: The main sources of oxygen replenishment in river water are absorption from the atmosphere and photosynthesis of aquatic plants and algae. The term reaeration is used to describe the atmospheric absorption of oxygen. The rate of reaeration is proportional to the dissolvedoxygen deciency. The rate of reaeration, rR , can be expressed as
rR = k2 (Cs CO2 )

(5.5)

where

rR k2 Cs CO2

= = = =

rate of reaeration, g m3 d1 reaeration constant, d1 dissolved-oxygen saturation concentration, g m3 dissolved-oxygen concentration, g m3 .

The re-aeration constant can be estimated from the characteristics of the stream and an appropriate formula such as the one suggested by OConnor and Dobbins: p 294 O2 U k2 = , (5.6) h3/2 where
O2 U h

= = =

kinematic viscosity of oxygen (temperature dependent), m2 s1 , stream ow velocity, m s1 , and average depth of ow, m.

If, for example, we have a stream with U = 1 m s1 and h = 2 m, and noting that O2 = 1.5 105 m2 s1 at 20o C, k2 = 0.40 d1 . The values of O2 at temperatures other than 20o C are given by the relationship
O2 (T ) = O2 (20) 1.037T 20 .

(5.7)

The Oxygen-Sag Model: Here we ignore the effects of dispersion until later. In this way we can interchange between time and space and assume that space x and time t are simply related by the mean velocity of ow in the river. The deoxygenation and reoxygenation processes will now be considered in an oxygen mass balance equation, which we now develop, assuming rather unrealistically, that the river and waste are completely mixed where the waste enters the river and that wastes discharged into the river are distributed uniformly over the cross-section. In general this assumption is only true at distances well downstream of the waste discharge input point. Initial Mixing: the initial concentration, C0 , of a constituent in the river-waste mixture at x = 0 is given by Qr Cr + qw Cw C0 = (5.8) Qr + qw

49

River hydraulics

John Fenton

Figure 5-2. Longitudinal section of river of cross-sectional area A with oxygen uxes crossing boundary of control surface. Deoxygenation is internal.

t
( x, t )

x
x /U
1

U
( 0, t 0 )

Figure 5-3. Plot on (x, t) axes showing the characteristic (path) a parcel of uid follows

where

C0 Cr Cw Qr qw

= = = = =

initial concentration of constituent at point of discharge, g m3 concentration of constituent in river before mixing, g m3 concentration of constituent in wastewater, g m3 . river discharge, m3 s1 wastewater discharge, m3 s1

Oxygen Balance: Consider, steady uniform one-dimensional ow in a prismatic river channel, a wastewater input discharge with complete mixing within the channel cross-section, and no longitudinal dispersion, as shown in Figure 5-2. Consider now the mass balance (mass conservation) of the oxygen in the river for an incremental volume given by A x as shown in the gure:
CO2 CO2 A x = QCO2 (QCO2 + Q x) + rD A x + rR A x. | {z } | {z } {z | t {z } |Inow} | {z x } Deoxygenation Reoxygenation
Accumulation Outow

(5.9)

Before substituting rD and rR and attempting to solve, it is very helpful to consider the trajectory or path along which a parcel of mixture takes and then to solve the partial differential equation along that trajectory. Figure 5-3 shows such a path on (x, t) axes, where the parcel sets out at a time t0 . If the mean velocity U = Q/A in the stream is constant, the trajectory is a straight line, starting at t = t0 , with gradient 1/U , and the equation of the trajectory is
t = t0 + x . U

(5.10)

At time t, the time over which deoxygenation has taken place is t t0 ,and so we modify equation (5.4), 50

River hydraulics

John Fenton

replacing t by t t0 :

rD = k1 BODu ek1 (tt0 ) .

(5.11)

Substituting equations (5.11) and (5.5), dividing through by Ax and re-arranging, we have the partial differential equation for the dissolved oxygen concentration of a parcel of mixture which started at x = 0 at t = t0 : CO2 CO2 +U + k2 CO2 = k1 BODu ek1 (tt0 ) + k2 Cs . (5.12) t x Now, as we want to consider x and t related by the equation of the trajectory, the two are no longer independent. We enter a general discussion for combinations of derivatives as in equation (5.12), which occur throughout uid mechanics. The material or advective derivative: At a xed point in space the rate of change of a quantity such as temperature or even the velocity vector, is /t. However, even in a steady ow eld uid particles experience a different apparent rate of change by moving to a position where has a different value. If a uid element at x at time t is at x + x at time t + t, we can write the Taylor expansion for the value of at the new point:
(x + x, t + t) = (x, t) + t + x + higher order terms, t x

hence the rate of change of at the uid particle, denoted by D/Dt is


D Dt = = lim (x + x, t + t) (x, t) lim x = + t 0 t t t 0 t x +U . t x

The quantity D/dt is the Advective Material Derivative, the time rate of change experienced by a uid particle. The rst term is the temporal rate of change, which is zero in steady ow, while the second is the advective contribution, which exists because the particle moves to a point where the velocity is different it is being transported through a eld where there is a gradient in at a velocity U . In a uid ow the rate of change at a particle is different from that at a point. Solution: Immediately we recognise that we can write equation (5.12) as
DCO2 + k2 CO2 = k1 BODu ek1 (tt0 ) + k2 Cs , Dt

(5.13)

an ordinary differential equation for CO2 . For solution we will impose the initial condition that CO2 (0, t) = C0 (t0 ). Equation (5.13) is in a form suitable for use of an integrating factor, such that if we multiply both left and right sides by the quantity I = exp k2 (t t0 ), then it can be written: D CO2 ek2 (tt0 ) = k1 BODu e(k2 k1 )(tt0 ) + k2 Cs ek2 (tt0 ) . Dt

Integrating both sides with respect to t (for mathematical solution the advective derivative has the same properties as an ordinary derivative):
CO2 ek2 (tt0 ) = k1 BODu e(k2 k1 )(tt0 ) + Cs ek2 (tt0 ) + B, k2 k1 k1 BODu + Cs + B, k2 k1

where B is a constant of integration. Satisfying the initial condition gives


C0 (t0 ) =

51

River hydraulics

John Fenton

In this form it is still applicable to the trajectory described above. We can write it such that it is applicable to any (meaningful) value of x and t by substituting t t0 = x/U and C0 (t0 ) = C0 (t x/U ), giving k1 CO2 (x, t) = C0 (t x/U ) ek2 x/U + Cs 1 ek2 x/U + BODu ek2 x/U ek1 x/U . k2 k1 (5.15) This is a generalised version of the Streeter-Phelps oxygen sag equation. In this form the exponential functions on the right are functions of the independent variable x, and time t appears explicitly in only one place. To obtain the value of the concentration at a particular place x and time t one evaluates the functions of x and computes the corresponding initial value of concentration which was introduced at a time t x/U , i.e. prior to time t by an amount x/U , the travel time taken to get to the point (x, t) required. Equation (5.15) seems to have advantages over conventional presentations of the Streeter-Phelps oxygen sag equation, which is presented in textbooks only for the case of constant C0 and with time t as an independent variable even in the case where the whole problem is steady. The oxygen-sag equation, with its two exponential decay rates, is as shown in Figure 5-4. Biological decomposition begins immediately following waste input and uses oxygen from the river with a time scale of k1 and a space scale of k1 /U . Atmospheric reaeration is proportional to the dissolved-oxygen decit and therefore increases with increasing decit giving a gradual mobilisation of the reaeration process with time and space scales of k2 and k2 /U . Eventually a critical point xc is reached when waste decomposition rate equals the rate of atmospheric re-aeration. Downstream of this point the rate of re-aeration exceeds the rate of decomposition and the dissolved oxygen level begins to increase until nally saturation conditions are reached. This model is a simple but commonly used one for modelling oxygen resources in a river. It neglects the effects of algae and sludge deposits on the oxygen balance, as well as, more importantly, all effects of dispersion due to natural processes including turbulence in the stream. Later we will spend some time examining these.

which can be solved for B and substituted into the solution to give k1 CO2 = C0 (t0 ) ek2 (tt0 ) + Cs 1 ek2 (tt0 ) + BODu ek2 (tt0 ) ek1 (tt0 ) . (5.14) k2 k1

Figure 5-4. Characteristic oxygen-sag curve obtained using the Streeter-Phelps equation

52

River hydraulics

John Fenton

Critical dissolved oxygen decit: The critical or maximum dissolved-oxygen decit Dc will occur when dCO2 /dt = 0. This can be found from equation (5.13) with this condition, leading to
CO2 = Cs k1 k1 BODu ek1 (tt0 ) = Cs BODu ek1 x/U . k2 k2

To nd when this occurs we would have to differentiate the solution, equation (5.15) with respect to x. For a constant C0 we obtain k2 k2 (Cs C0 ) U x= log 1 . k2 k1 k1 k1 BODu Exercise: Verify that equation (5.15) satises (a) the required boundary condition CO2 (0, t) = C0 (t), and (b) the original differential equation (5.12) (after substituting t t0 = x/U on the right side). Hint: you will need the result from the theory of partial differentiation that
0 C0 (t x/U ) = C0 (t x/U ) and t 1 0 C0 (t x/U ) = C0 (t x/U ), x U

0 where C0 (t x/U ) indicates the ordinary derivative of the function with respect to the argument shown.

5.9 Salinity in rivers


Salinity refers to the concentration of total dissolved solids (TDS) in water and is perhaps historically the most important water quality characteristic because of its relationship to agriculture. It is a gross chemical characteristic and can be expressed either in terms of:
Concentration in g m3 mg l1 ppm (parts per million) or Electrical Conductivity of the water in MicroSiemens cm1 at 25o C, designated by S cm1 and known as EC units.

The relationship between EC and TDS is given approximately by 1 S cm1 or EC unit 0.64 mg l1 of total dissolved solids. 5.9.1 Tolerance Levels Salinity limits have been established for various uses of river water. Although there is no sharp division between acceptable and unacceptable salinities, it is widely recognised that economic losses increase with increasing salinity. According to the former Rural Water Commission, Victorian waters can be classied as follows: Class 1: TDS 0-175 ppm low salinity water which can be used with most crops on most soils, with little likelihood that a salinity problem will develop. Class 2: TDS 175-500 ppm medium salinity water which can be used with plants of medium salt tolerance without adopting special practices for salinity control. Class 3: TDS 500-1500 ppm high salinity water which cannot be used on soils with restricted drainage, and even with adequate drainage, the salt tolerance of any plants to be irrigated must be taken into account. Class 4: TDS 1500-3,500 ppm very high salinity water not suitable for irrigation under ordinary conditions. Class 5: TDS above 3,500 ppm extremely high salinity water to be used in emergencies.

These TDS levels are only approximate and concentration of particular salts within the total dissolved 53

River hydraulics

John Fenton

solids can be important. 5.9.2 Human Consumption The suitability of water for human consumption depends not only on the salinity level but on the presence of organisms indicative of harmful pollution. The World Health Organisation has published international standards for drinking water which recommend an upper limit of salinity of 1500 ppm. The limit in general use in Victoria is 2000 ppm if no other water is available. However the desirable maximum salinity is 835 EC units (530 g l1 or ppm) according to the Engineering and Water Supply Department of South Australia. 5.9.3 Septic Tanks Bacterial action in septic tanks can be maintained at very high levels of salinity. 5.9.4 Sources of Salinity in Rivers The salinity of river water can derive from a number of sources.
Mineral salts are released by rock weathering, collect in the soil, and are washed into streams by natural runoff. Rainfall can carry small concentrations of oceanic salts. Saline groundwater can be seeping into a river. Where a saline water table rises to within capillary range of the surface accumulations of salt will occur in a river catchment and will be eventually washed into the river system. The rise in the water table could be the result of excessive land clearance or unsatisfactory irrigation practices. The salinity level in a river increases in the downstream direction as tributaries add their contributions of salt. although high salinity levels occur naturally in some rivers, particularly during low ow periods, these levels are increased wherever human activities in the river catchments have resulted in rising water table levels.

5.9.5 The Murray River The Murray River has been chosen to illustrate the variation of salinity along a river since it is our major river. It supplies water for extensive irrigation areas and water to augment the water supplies of most South Australian towns and cities. Figure 5-5 shows two longitudinal proles of salinity along the Murray River from Hume Dam to the river mouth in South Australia. Both proles recorded in the early 1970s, indicate that salinity is low until the Loddon River conuence is reached, when highly saline water from Barr Creek enters the Murray. Barr Creek is the main drain of a surface drainage network serving about 125,000 ha of saltaffected farmland in the Kerang region. It was the biggest point source of salt along the river and caused a marked salinity jump. (In recent years measures have been taken to restrict the highly saline ows of Barr Creek entering the Murray). The next major input is from the Wakool River (whose main source of salt is from groundwater seepage). Its effect can be masked by signicant ows from the Murray which divert down the Edward River and dilute Wakool ows before they enter the Murray. Further downstream, the inow of the Murrumbidgee River provides a signicant diluting effect. Near Merbein, and in the 20km between Mildura and Merbein, major salt inputs occur as a result of seepage of highly saline groundwater. Inputs of highly saline groundwater are the dominant source of salt input in the lower reaches of the Murray River. The salinity of the groundwater derives from the late Tertiary Era when a gulf of sea extended eastwards into the Murray Basin as far as Swan Hill, covering much of todays Mallee Region. When the sea began to retreat about one million years ago it left soil and groundwater salty. 54

River hydraulics

John Fenton

Figure 5-5. Longitudinal salinity proles in the Murray River, after Collett (1978)

The longitudinal proles shown in the gure illustrate salinity levels which are of great concern to South Australia. 5.9.6 Measures for Salinity Mitigation in Rivers Three important approaches can be identied:
Restricting the entry of saline water to the river, particularly during critical periods. This could involve for example restricting the entry of highly saline water from tributaries (as in the case of Barr Creek in the Murray River example) or lowering the water table levels near points where saline groundwater enters a river. Providing ushing ows of good quality water to dilute the concentration of dissolved solids in a river during critical periods. One important effect of river regulation is the improved salinity situation in low ow periods, particularly in drought years. Improving water management in irrigation districts, particularly with respect to drainage and control of water table levels. This aspect is outside the scope of these notes.

5.9.7 Modelling of Salt Transport in Rivers The impact of salinity mitigation schemes and river operating policies for reducing salinity can be examined using an appropriate numerical model describing salt transport down the river system. Since salt transport is totally dependent on the movement of water through the system, it is rst necessary to predict the velocity eld before solving for the distribution of salt concentration. a one-dimensional approach is usually adopted for the model. One approach is to start from the partial differential equations governing the ow and the distribution of salt concentration, and then to solve these numerically by converting them into algebraic equations for discrete points or elements. A nite-difference approach is usually adopted. Care must be exercised in 55

River hydraulics

John Fenton

the choice of an appropriate discretisation scheme, since numerical methods can introduce numerical diffusion which may distort the solution. The basic one-dimensional ow equations are the Saint-Venant or Long Wave equations, which we will now develop. These are widely used for studies of the movement of water in waterways, including ows and oods in rivers, and the delivery of water in irrigation channel systems. 5.9.8 Further Reading The articles by Collett (1978) and Keiller & Close (1985) are interesting.

6. Turbulent diffusion and dispersion


We consider ows which are reasonably steady (albeit turbulent) and have uniform geometry. For these cases, we can use a Fickian-type of law for diffusion due to turbulence. Initially as an introduction, let us consider a local velocity eld which is turbulent. We will only consider the x component, so that we write
u = u + u0 ,

where u is the time-mean velocity at a point, and u0 is the uctuating velocity, such that the time mean of u0 is zero, u0 = 0. Similarly, we write an expression for concentration of a pollutant
c = c + c0 ,

and we want to calculate the ux of the pollutant, uc uc = u + u0 c + c0 = u c + u c0 + c u0 + u0 c0 . However, we do not want to study the uctuating values of this, and so we take the time mean:
uc = u c + uc0 + cu0 + u0 c0 = u c + u c0 + c u0 + u0 c0 = u c + u0 c0 ,

and so we have the possibly-surprising result that the mean of uc is not the mean of u times the mean of c, but contains the extra turbulent transport ux u0 c0 . Of course, there are other such quantities in the equations of hydraulics notably quantities such as u02 , u0 v 0 , u0 w0 , etc. in the equations of motion. Quantities such as the latter three are known as the Reynolds stresses. One of the most important problems in uid mechanics is to model these turbulent quantities in terms of the mean quantities of the ow. The turbulent transport uxes such as u0 c0 have several attributes:
They range in size from uctuations the size of the ow geometry down to the smallest turbulent scale. They are random individual uctuations are not repeatable, but the statistical characteristics of those uctuations are. The random nature of the ow results in intermittent but large spatial gradients in variables as packets of uid with quite different characteristics are brought into contact by the turbulence. This enhances molecular diffusion. The continuous introduction of packets of high and low concentration by the turbulence will result in an overall dilution of zones of high concentration via turbulent mixing. This is turbulent diffusion, which is rather more important than molecular diffusion in river ows.

The energy to create turbulence is typically put into the system at scales the size of the overall geometry. 56

River hydraulics

John Fenton

Viscosity is the agent ultimately acting so as to dissipate the turbulent energy in the form of heat. It operates on the smallest possible scales of the ow eld at a molecular level. In between the creation and dissipation turbulent scales there is an orderly cascade of turbulent energy by which energy is nonlinearly transmitted from the large to small turbulent uctuations or scales by a continuously decreasing sequence of turbulent eddies. This has been tersely described by L. F. Richardson, the father of computational uid mechanics, paraphrasing Jonathan Swift: Big whirls have little whirls that feed on their velocity, And little whirls have lesser whirls and so on to viscosity - in the molecular sense.

6.1 Diffusion and dispersion in waterways


With the exception of free jets such as smoke or exhaust stacks in the atmosphere, most ows encounter and are affected by boundaries. These cause the ow eld and its turbulence characteristics to be different in all three co-ordinate directions. It is anticipated that the eddy diffusivities will be different in each direction. The main objective here is to be able to predict the eddy diffusivities in turbulent channel ow and the importance of their differences in magnitude on turbulent mixing in the channel. 6.1.1 Eddy diffusivities Gradients of momentum and transport are sharper in the vertical and the transverse direction. An approximation obtained by Elder for the mean vertical eddy diffusivity z is
z = 0.067 u D,

where D is the depth, and u is the shear velocity, r 0 p gDS0 , u = where 0 is the mean boundary shear stress. For the transverse eddy diffusivity y Fischer et al. (1979) suggest
z 0.15 u D,

with a similar result for the longitudinal diffusivity x :


x > 0.15 u D.

The coefcients vary, depending on the channel or estuary type.

6.2 Dispersion
An important feature of turbulent diffusion in waterways is that the velocity distribution is not uniform across the stream. This has the ability to separate individual particles much more quickly than if just turbulent processes were at work, and for the effective diffusivity to be considerably greater. This is shown in Figure 6-1. Part (a) shows a hypothetical stream where the velocity is constant across the section. A puff of pollution introduced uniformly across the section is then carried downstream, and turbulent diffusion plays the role that we see, whereby the concentration distribution is gradually diminished and widened, according to the principles we have already studied. The top gure shows an instantaneous snapshot of the pollution cloud, and in the next part of the gure the mean concentration distribution across the stream is plotted, which will closely approximate a Gaussian. In Part (b) of the gure, however, we see what happens when there is a realistic velocity prole. Some particles near the side do not travel as far as those in the middle, and the cloud of pollution becomes increasingly distorted, and the extent of the cloud can quickly become large. Now, taking the mean across the stream we see the results in the bottom part of the gure, where the mean concentration is now much lower, as is the lateral extent 57

River hydraulics

John Fenton

(a) Hypothetical uniform velocity profile

Plan

(b) Realistic velocity profile

Plan

Figure 6-1. The role of the velocity distribution in enhancing dispersion

of the cloud. The mean diffusivity has been considerably increased by the velocity prole. This phenomenon was rst described by G. I. Taylor in 1953. This artifact of spatial averaging is known as dispersion. The basic operation involved is to obtain a partial differential equation which is derived in terms of area-averaged variables. By considering the sort of control volume we used for the derivation of the oxygen-sag equation, Figure 5-2. We can easily show, generalising our original analysis to nonuniform distributions over the section (but with no de-oxygenation or reaeration), that Z Z c dA + cu dA = 0. t x
A A

Now, in a similar sense to what we did for time-uctuating quantities we let


c = e+ c , c
00 00

where e is the spatial average over a section, and c is the amount by which the concentration varies c from that average. Writing a similar expression for u, and substituting, we obtain Z Z 00 00 c dA = eA, and c cu dA = euA + cg A, ce u
A A

giving

g (eA) + c (euA) + ce c00 u00 A = 0. t x x However, u is simply the mean velocity as we have already used it u = U = Q/A,and if we expand the e e terms in this equation and subtract the mass conservation equation A + (eA) = 0, u t x

then we obtain

e c e c g +u e + c00 u00 = 0, t x x

58

River hydraulics

John Fenton

where we have neglected an area derivative in the uctuating component term (area changes relatively slowly).
00 00 Taylor postulated (see e.g. Fischer et al. 1979) that cg can be modelled as u

where T is the dispersion coefcient. Substituting, and neglecting derivatives of this quantity, we obtain
e c 2e c e c +U = T 2 , t x x

e c 00 00 cg = T , u x

the advection-diffusion equation, but where the coefcient of diffusion is the dispersion coefcient T . Fischer et al. (1979) and French (1985) give a detailed method for estimating T . There is a simple empirical formula:
U 2T 2 , u D where T is the top width, as used previously. This expression is correct to within a factor of four, which is considered acceptable ... For a narrow channel, a typical value (p423 of Streeter et al. 1998) for a mean ow velocity of 0.1 m s1 is 2.75 m2 s1 , and for a wide channel a typical value is 275 m2 s1 . T = 0.011

6.3 Non-dimensionalisation Pclet number and Reynolds number viscosity as diffusion


Consider a solution of the advection-diffusion equation over a physical length L. If we non-dimensionalise such that x = x/L, then the computational domain becomes (0, 1), which is convenient. If we also non-dimensionalise time with respect to t = t L/U , then we obtain
U U 2 + = 2 2, L t L x L x

and multiplying through we obtain


2 + = Pe 2, t x x

where P e =

. UL

The Pclet number P e is a dimensionless number which expresses the relative importance of diffusion to advection. It has appeared here as a result of a non-dimensionalisation. Clearly we could use this to express rather more concisely solutions to the advection-diffusion equation, as it is the only parameter governing the solution.

7. Sediment motion
The grains forming the boundary of an alluvial stream have a nite weight and nite resisting ability, including cohesion and coefcient of friction. They can be brought into motion if the forces due to uid motion acting on a sediment particle are greater than the resisting forces. Often this is expressed in terms of disturbing and resisting stresses on the bed of the stream. If the shear stress acting at a point on the ow boundary is greater than a certain critical value cr then grains will be removed from that region, and the bed is said to scour there. We introduce the concept of a relative tractive force at a point, / cr . If this is slightly greater than 1, only the grains forming the uppermost layer of the ow boundary can be detached and transported. If / cr is greater than 1, but less than a certain amount, then grains are transported by deterministic jumps in the neighbourhood of the bed. This mode of grain transport is referred to as bed-load. If the ratio is large, then grains will be entrained into the ow and will be carried downstream by turbulence. This transport mechanism is known as suspended-load. The total transport 59

River hydraulics

John Fenton

rate is the sum of the two. The simultaneous motion of the transporting uid and the transported sediment is a form of two-phase ow. We can write all the variables which should dominate the problem of the removal and transport of particles:
s g h

Density of water Density of solid particles Kinematic viscosity of water Diameter of grain Gravitational acceleration Depth of ow Shear stress of water on bed

ML3 ML3 L2 T1 L LT2 L ML1 T2

As we have 7 such quantities and 3 fundamental dimensions involved, there are 4 dimensionless numbers which can characterise the problem. In fact, it is convenient to replace g by g 0 = g (s / 1), the apparent submerged gravitational acceleration of the particles, and to replace by the shear velocity p u = /. Convenient dimensionless variables, partly found from physical considerations, which occur are
=

u2 = , roughly the ratio of the shear force on a particle to its submerged weight 0 g (s ) g u , roughly the ratio of uid inertia forces to viscous forces on the grain R = G = s , the specic gravity of the bed material, and , the ratio of grain size to water depth h

Two important quantities here are R which is the grain Reynolds number, and the Shields parameter, which can be thought of as a dimensionless stress.

7.1 Incipient motion


In the 1930s Shields conducted a number of experiments in Berlin and found that there was a narrow band of demarcation between motion and no motion of bed particles, corresponding to incipient motion. He represented these on a gure of versus R . A slight problem with this is that the uid velocity (in the form of shear velocity) occurs in both quantities. It is more reasonable to introduce the dimensionless grain size (see p7 of Yalin & Ferreira da Silva 2001): 2 1/3 0 1/3 R g = = . 2 Here we consider what means. If we take a common value of G = 2.65, plus g = 9.8 m s2 , = 106 m2 s1 (for 20 C), then we obtain 25000 in units of metres. If is specied in terms of millimetres then we have 25 , and so for a range of particle sizes we have
( mm) 0.1 0.004 1 0.04 10 0.4 100 4 1000 40

Above the line, for larger values of (and hence larger velocities or smaller and lighter grains), particles 60

Figure 7-1 shows a representation of Shields results, using for the abcissa. Instead of the experimental results we use a formula by Yalin which is an approximation to the results for incipient motion, giving the critical value cr : 2 cr = 0.13 0.392 e0.015 + 0.045 1 e0.068 . (7.1)

River hydraulics

John Fenton

Yalins approximation to Shields data, eqn (7.1) Bagnolds conjecture for random beds All beds random on the scale of the particles Motion

Dimensionless 0.1 shear stress No motion

Beds at in laboratories

Beds random in nature 0.01 0.1 1 10 Dimensionless grain size 100 1000

Figure 7-1. Incipient motion diagram

will be entrained into the ow. Below the line, particles should be stable. For small particles there appears to be a linear relationship (on these log-log axes), while for large particles the critical shear stress, based on Shields laboratory experiments, approaches a constant value of about 0.045. In between there is a dip in the curve, with a minimum at about = 14, corresponding to a grain size of 0.35 mm, about a ne sand. Bagnold (personal communication) suggested that there was probably no uid mechanical reason for that, but that there is an implicit scale effect in the diagram, an articial geometric effect, and suggested that the Shields diagram has been widely misinterpreted. He suggested that for experiments with small particles, while the overall bed may have been attened, individual small grains may sit on top of others and may project into the ow, so that the assemblage is random on a small scale. For large particles (gravel, boulders, etc.) in nature, they too are free to project into the ow, however in the experiments which determined the Shields diagram, the bed was made at by levelling the tops of the large particles. Hence, there is an articial scale effect, and if one were only to consider random beds of particles which are free to project into the ow above their immediate neighbours, while the bed might level on a scale much larger than the particles. Fenton & Abbott (1977) followed Bagnolds suggestion and examined the effect of protrusion of particles into the stream. Although they did not obtain denitive results, they were able to recommend that for large particles the value of cr was more like 0.01 than 0.045, which seems to be an important difference, the factor of 1/4 requiring a uid velocity for entrainment into the ow of randomly-placed particles to be about half that of the sheltered case. A curve representing Bagnolds hypothesis, as partly borne out by the experiments, is shown on Figure 7-1.

7.2 Relationships for uvial quantities


Now we consider some simple relations to relate the above results to real physical quantities. The shear velocity u is a very convenient quantity indeed. If we consider the steady uniform ow in a channel, then the component of gravity force down the channel on a slice of length x is gA x S0 . However the shear force resisting the gravity force is P x. Equating the two we obtain
= g A S0 . P

61

River hydraulics

John Fenton

In this work it is sensible only to consider the wide channel case, such that A/P = h, the depth, giving
= gh S0 ,

or in terms of the shear velocity:


u =

and so in terms of the dimensionless stress:


=

p = gh S0 ,

gh S0 S0 h u2 = 0 = , 0 g g (G 1)

giving a result which can be used in association with the Shields diagram or equation (3.15) to give the slope, depth, or particle size at which entrainment will occur.

7.3 Dimensional similitude


In experiments with sediment transport, as in other areas of uid mechanics, it is desirable to have the same dimensionless numbers governing both experimental and full-scale situations. In this case we would like the dimensionless particle size AND the dimensionless shear stress each to have the same values in both model and full scale. Using the subscript m for model and no subscript for the full scale situation, we then should have 0 1/3 0 1/3 gm g m = such that m = , 2 2 but as gravitational acceleration g and viscosity can be assumed to be the same in each, we can write
m (Gm 1)1/3 = (G 1)1/3 .

This means that


m =

Obviously if we use the same material for the model as for the full scale, Gm = G, and we have m = . But if we use a lighter material in the model we will have to have m > . Also we require the same dimensionless shear stress for the model and the full scale:
S0m hm S0 h . = (Gm 1) m (G 1)

G1 Gm 1

1/3

(7.2)

However, if we eliminate particle size using the previous relation (7.2) then we obtain
(Gm 1)2/3 S0m hm = . S0 h (G 1)2/3

We can see that if we use the same material, the right side is unity and we have S0m hm = S0 h, so that if we use a model scale of, say, hm /h = 1/10, then we have S0m = 10 S0 , which is fortunate, as shallow slopes are difcult to arrange in a laboratory. On the other hand, if we use a rather lighter material in the laboratory, as is often the case, then we will require S0m hm < S0 h, and possibly end up requiring a laboratory slope similar to that in the eld, which might be very difcult to arrange.

7.4 Bed forms


In most practical situations, sediments behave as non-cohesive materials, and the uid ow can distort the bed into various shapes. The interaction process is complex. At low velocities the bed does not move. With increasing ow velocity the inception of movement occurs. The basic bed forms encountered are 62

River hydraulics

John Fenton

ripples (usually of heights less than 0.1 m), dunes, at bed, standing waves, and antidunes. At high velocities chutes and step-pools may form. Typical bed forms are summarised in Figure 7-2 below.

Figure 7-2. Bedforms and the Froude numbers at which they occur (after Richardson and Simons)

7.5 Mechanisms of sediment motion


The amount of sediment passing through a given stream cross-section is a critical factor in
determining the amount of sediment deposited in a downstream reservoir, and thus the useful life of a dam or project investigating the stability of the streambed and stream banks inuencing the water quality to be used for irrigation, water supply, and recreational considerations affecting the navigable depth of a stream determining the water level during oods studying the potential effects on stream ecology such as sheries as well as other biological species investigating the movement of pollutants attached to sediment particles

7.5.1 Bed load the movement of sediment particles along the streambed in the process of rolling, sliding, and/or saltation, when the ow turbulence picks up a particle from the bed and they then fall back to the bed, somewhere further downstream. There are no readily-accepted formulae for the movement of gravel, as the process of armouring may or may not take place, which gives widely variable results. For sand, however, there are several popular equations. Einsteins bed load equation: 63

River hydraulics

John Fenton

7.5.2 Suspended load the movement of sediment particles that are supported by the turbulent motion in the streamow. 7.5.3 Bed-load rate of transport Bagnolds formula The volumetric rate of transport qsb per unit width is given by
qsb = ub ( cr ) , (s ) g

where is a function of , and ub is the ow velocity in the vicinity of the bed. In the case of a rough turbulent ow, 0.5. This formula is preferred, as it is simple, as accurate as any, and reects the meaning of the bed-load rate. However the calculation of ub seems a little arbitrary.

64

River hydraulics

John Fenton

Appendix A. On diffusion and von Neumann stability analyses


A.1 The nature of diffusion
A.1.1 A discrete physical analogy
n 0 1 2 3 4 m= -4 -3 -2 1/4 1/4 -1 1/2 3/8 0 1 1/2 3/8 1 1/2 3/8 2 1/4 1/4 3 4

1/16

1/8

1/8

1/16

Figure A-1. Array of pins showing probabilities with which a ball will pass through a particular gap

The process of diffusion occurs because of a continuous process of random particle movements. We will now model that, but instead of a multi-dimensional medium where random movements of any magnitude can occur, we will consider a one-dimensional medium where a single particle is free to make a series of discrete movements of the same magnitude, either to left or right. Consider a ball dropped between the two pins at the apex of the pyramid of pins shown in Figure A-1. The probability that it will pass between the rst two pins is 1. It will hit the pin beneath it and will roll to left or right, with an equal probability. The next lower set of pins in the discrete physical analogy is equivalent to the next time stage in the real physical situation. The relative probabilities of dropping down each gap is (1, 1), with the actual probabilities obtained by dividing by the sum of those relative probabilities, giving (1/2, 1/2). Similarly at the next level of pins, the probabilities are (1, 2, 1), giving (1/4, 1/2, 1/4), and at the next level, (1, 3, 3, 1), or relatively, (1/8, 3/8, 3/8, 1/8) and so on, familiar to anybody who has expanded, say, (a + b)3 . The pattern is obvious in fact those probabilities are given by the binomial coefcients, which we can express as
1 n! (m, n) = n+m nm n , ! !2 2 2

(A-1)

for the nth row of pins, starting from n = 0 at the rst row where all balls pass through the same gap, and the mth gap where m goes from n to n in steps of 2, avoiding the places occupied by the pins. A remarkable result is now used, Stirlings approximation for the factorial function: k! kk ek 2k, for large k. Using this it can be shown Borg & Dienes (1988) that equation (A-1) can be approximated by r 2 m2 /2n e (m, n) , n

(A-2)

and we see that at the nth row of pins, the shape of the probability distribution is approximated by 2 the Gaussian function em /2n which occurs throughout statistics. This has a characteristic bell-shape. 2 Note that variation with m (the horizontal passage position or co-ordinate) varies like em , and the coefcient = 1/2n, such that it decreases with n, showing how the lateral extent increases with n (time). Figure A-2 shows a comparison between the binomial distribution, equation (A-1) and the Gaussian distribution (A-2). We now wish to relate this to real diffusion processes and quantify the physical situation rather more. We suppose that a particle of a diffusing substance can either move to left or right a distance l in a time . Table A-1 shows the calculation of the mean square displacement obtained from the discrete model we considered above. The last column demonstrates the empirical relation for the mean square displacement:
2 = 1 X 2 l2 =t , N

65

River hydraulics

John Fenton

0.5

Binomial Approximation

0.4

Probability

0.3

0.2

0.1

0 -10

-8

-6

-4

-2

0 m

10

Figure A-2. Probability distributions of particle positions at successive levels

Time t 0 2 3

Possibilities for 0 l 1 0 2, 2l 1 l 3, 3l 1

No. of possibilities N 1 21 22 24

2 0 2l2 8l2 24l2

2 =

1 2 N

0 l2 2l2 3l2

Table A-1. Mean square displacement 2 for random walk as a function of the diffusion time

showing that it is proportional to time (n in the above analogy). If we set


l2 = , 2

and call this quantity (of dimension L2 T1 ) the coefcient of diffusion, then we obtain
2 = 2t,

similar to an expression obtained by Einstein in his paper on Brownian motion in 1905. A.1.2 Ficks law

Figure A-3. Diffusion model with more particles crossing the plane at x = l/2 to the right than to the left

66

River hydraulics

John Fenton

Now we obtain the basic law governing diffusion. Instead of one particle at each of the different physical locations we postulated above, we consider a different number of particles at each. If a particle moves a displacement l in a characteristic time , the mean velocity of motion is l/ . There are probabilities of 1/2 that it will go either way. If the number of particles is n0 at x = 0, then the total ux of particles (number of particles times velocity) which move from x = 0 to x = l is
1 l J+ = n0 . 2

Similarly, if there are n1 particles at x = l, then the ux of particles in the other direction is
1 l J = n1 . 2

The total ux in the +x direction at x = l/2 is


J = J+ + J = (n1 n0 ) l2 . l 2

We have suggested that the quantity l2 /2 is the diffusion coefcient, a constant for a particular medium and diffusing molecules. If we denote the difference in the concentration of particles by c = n1 n0 over the distance x = l, we obtain c . J = x In the limit x 0 we obtain Ficks law for the diffusion ux:
J = c . x

Hence we see that the ux of molecules is proportional to the gradient of the concentration but that this does not come from any special physical property of the concentration, but merely because where there are many particles there will be more travelling away from that region than are travelling to that region. This is the essence of the diffusion process. In Brownian motion and diffusion in solids the random motion comes from the molecular natures of the constituents. In waterways the random motion imparted to individual elements is due to the turbulence in the water. A.1.3 Exact solutions of the diffusion equation The diffusion equation has a general solution which is exact for an innite domain (, +): 1 x2 . = (A-3) exp 4t 4t The factor 4 is there such that the total integral of this solution between and + is 1. The solution at time t 0 is an innitely large and innitesimally narrow concentration distribution. Thus the solution corresponds to the concentration distribution due to a point source of pollution. Solutions for successive times are shown in Figure A-4, which shows how the solution extends out to as time progresses. Another well-known analytical method for solution of the diffusion equation is the use of Fourier series, where the medium is of nite length. It can be shown by the method of separation of variables that a single term in the series which satises the diffusion equation is
= exp(j 2 k2 t) sin jkx,

where k = 2/L, where L is the length of the medium, and j is an integer which goes from 1 to . It is clear that the shorter components (larger j ) decay very much more quickly than the longer ones. With this, we have the familiar behaviour that discontinuities are smoothed out very quickly. Consider the situation shown in Figure A-5, which shows a discontinuity between two regions where the concentration is initially different. A solution was obtained in the form of a Fourier series with 67

River hydraulics

John Fenton

Figure A-4. General solution for t = 1/4, 1/2, 1, 2, 4, 8

coefcients exponentially decaying with time. The total length of the domain is L. Results are shown are for the initial condition t/L2 = 0, then for t/L2 = 1/100, for 9 subsequent steps with a time interval of t/L2 = 1/20, and nally t/L2 , the straight line. It is clear how the diffusion works very quickly initially, but then proceeds more slowly. Notice how we can express our parameter non-dimensionally in this form.

Figure A-5. Diffusion of a discontinuity

68

River hydraulics

John Fenton

A.2 Examining stability by the Fourier series (von Neumanns) method


A.2.1 Pure advection We introduce a general method for examining the stability of approximations to differential equations. Suppose that the solution to the difference equation (4.2) can be written
(x, t) = A (t) eikx , where i = 1, and as eikx = cos kx + i sin kx, the variation in x is a single wave with wavelength L = 2/k. This is not as arbitrary as it appears at rst, as we can in theory represent any (periodic) variation in x as a Fourier series, and as we consider linear equations with constant coefcients only, we can just restrict ourselves to a single term in the Fourier series such as this one. We now want to work out what the solution for A(t) is, as would be obtained by the numerical method. Substituting into our FTCS computational scheme (4.2): (x, t + ) = (x, t) u ( (x + , t) (x , t)) , 2 u A (t) eik(x+) A (t) eik(x) , 2

which gives
A (t + ) eikx = A (t) eikx

and dividing through by A (t) eikx gives


A (t + ) A (t)

C +ik e eik 2 = 1 iC sin k = 1

We consider the magnitude of the amplication factor |A (t + ) /A (t)| (the factor by which the amplitude of the solution changes in a single time step) by multiplying by the complex conjugate to give: A (t + ) 2 A (t) = (1 iC sin k) (1 + iC sin k) = 1 + C 2 sin2 k. Now the criterion for stability is that the amplitude ratio should be less than or equal to one, that is, A (t + ) A (t) 1,

however this can never be, as C 2 sin2 k > 0, and we conclude that the scheme is unconditionally unstable for any values of space or time steps, as we experienced above. Methods such as this can be used to analyse rather more complicated schemes which do not have the same simple physical interpretation we saw in Figure 4-2. A.2.2 Von Neumann stability analysis of FTCS scheme with diffusion Let = A(t) eikx . Substituting into equation (4.3) and dividing through as before, we obtain A (t + ) = 1 + D eik 2 + e+ik A (t) = 1 2D (1 cos k)

This is a purely real quantity. Stability will be assured if (A(t + )/A(t))2 6 1. This gives the condition, after factorising:
D (1 cos k) (D (1 cos k) 1) 6 0.

69

River hydraulics

John Fenton

The rst factor is positive, the second factor is positive or zero for all k , (i.e. for all possible wavelengths given by k) so that for stability the last factor must be negative. That is,
D (1 cos k) 1 6 0,

giving
1 . 1 cos k Over all possible values of k , the minimum value of the right side is 1/2, giving the criterion D6 D= 1 2 6 2

for stability. This is the value usually quoted in Geomechanics applications. A.2.3 Exact solutions for combined advection-diffusion In fact, because of the simplicity of the effects of advection, this can often be included almost trivially. Consider the advection diffusion equation (3.15). If we introduce another advective co-ordinate X = x ut, such that it travels with the ow, we try the solution
(x, t) = (X, t).

We have, from the chain rule for partial differentiation,


t x = = X + = u , t X t t X 2 X 2 = , and = , X x x x2 x2

so that the advection equation becomes


2 = , t X 2 and we can see that a solution of the diffusion equation in X will be a solution of the advection-diffusion equation in x. This is all right if we do not have to impose any boundary conditions, so it can provide insight but may not be practically very useful. We can use the solution from equation (A-3) to provide the solution on (, +) of the advection-diffusion equation: ! (x ut)2 1 exp . = 4t 4t

This solution is plotted on Figure A-6, and shows the behaviour of solutions of the advection-diffusion equation. Note that the time levels for the curves are not equally spaced, but are as given in the caption to Figure A-4.

Figure A-6. Exact analytical solution of advection-diffusion equation

70

Vous aimerez peut-être aussi