尊敬的 微信汇率:1円 ≈ 0.046078 元 支付宝汇率:1円 ≈ 0.046168元 [退出登录]
SlideShare a Scribd company logo
April 18, 2007 
Open channel hydraulics 
John Fenton 
Abstract 
This course of 15 lectures provides an introduction to open channel hydraulics, the generic name for 
the study of flows in rivers, canals, and sewers, where the distinguishing characteristic is that the 
surface is unconfined. This means that the location of the surface is also part of the problem, and 
allows for the existence of waves – generally making things more interesting! 
At the conclusion of this subject students will understand the nature of flows and waves in open 
channels and be capable of solving a wide range of commonly encountered problems. 
Table of Contents 
References . . . . . . . . . . . . . . . . . . . . . . . 2 
1. Introduction . . . . . . . . . . . . . . . . . . . . . 3 
1.1 Types of channel flowtobestudied . . . . . . . . . . . . 4 
1.2 Properties of channel flow . . . . . . . . . . . . . . 5 
2. Conservation of energy in open channel flow . . . . . . . . . . . 9 
2.1 The head/elevation diagram and alternative depths of flow . . . . . 9 
2.2 Critical flow . . . . . . . . . . . . . . . . . . . 11 
2.3 TheFroudenumber . . . . . . . . . . . . . . . . 12 
2.4 Waterlevelchangesatlocaltransitionsinchannels . . . . . . . 13 
2.5 Somepracticalconsiderations . . . . . . . . . . . . . 15 
2.6 Critical flowasacontrol-broad-crestedweirs . . . . . . . . 17 
3. Conservation of momentum in open channel flow . . . . . . . . . 18 
3.1 Integralmomentumtheorem . . . . . . . . . . . . . . 18 
3.2 Flow under a sluice gate and the hydraulic jump . . . . . . . . 21 
3.3 The effects of streams on obstacles and obstacles on streams . . . . 24 
4. Uniform flowinprismaticchannels . . . . . . . . . . . . . . 29 
4.1 Features of uniform flow and relationships for uniform flow . . . . 29 
4.2 Computationofnormaldepth . . . . . . . . . . . . . 30 
4.3 Conveyance . . . . . . . . . . . . . . . . . . . 31 
5. Steady gradually-varied non-uniform flow . . . . . . . . . . . 32 
5.1 Derivation of the gradually-varied flowequation . . . . . . . . 32 
5.2 Properties of gradually-varied flow and the governing equation . . . 34 
5.3 Classification system for gradually-varied flows . . . . . . . . 34 
1
Open channel hydraulics John Fenton 
5.4 Somepracticalconsiderations . . . . . . . . . . . . . 35 
5.5 Numerical solution of the gradually-varied flowequation . . . . . 35 
5.6 Analyticalsolution . . . . . . . . . . . . . . . . . 40 
6. Unsteady flow . . . . . . . . . . . . . . . . . . . . 42 
6.1 Massconservationequation . . . . . . . . . . . . . . 42 
6.2 Momentum conservation equation – the low inertia approximation . . 43 
6.3 Diffusion routing and nature of wave propagation in waterways . . . 45 
7. Structures in open channels and flowmeasurement . . . . . . . . . 47 
7.1 Overshotgate-thesharp-crestedweir . . . . . . . . . . . 47 
7.2 Triangularweir . . . . . . . . . . . . . . . . . . 48 
7.3 Broad-crested weirs – critical flowasacontrol . . . . . . . . 48 
7.4 Freeoverfall . . . . . . . . . . . . . . . . . . . 49 
7.5 Undershotsluicegate . . . . . . . . . . . . . . . . 49 
7.6 Drownedundershotgate . . . . . . . . . . . . . . . 50 
7.7 DethridgeMeter . . . . . . . . . . . . . . . . . 50 
8. The measurement of flowinriversandcanals . . . . . . . . . . 50 
8.1 Methodswhichdonotusestructures . . . . . . . . . . . 50 
8.2 The hydraulics of a gauging station . . . . . . . . . . . . 53 
8.3 Ratingcurves . . . . . . . . . . . . . . . . . . 54 
9. Loose-boundary hydraulics . . . . . . . . . . . . . . . . 56 
9.1 Sedimenttransport . . . . . . . . . . . . . . . . . 56 
9.2 Incipientmotion . . . . . . . . . . . . . . . . . 57 
9.3 Turbulent flowinstreams . . . . . . . . . . . . . . . 58 
9.4 Dimensionalsimilitude . . . . . . . . . . . . . . . 58 
9.5 Bed-loadrateoftransport–Bagnold’sformula . . . . . . . . 59 
9.6 Bedforms . . . . . . . . . . . . . . . . . . . 59 
References 
Ackers, P., White, W. R., Perkins, J. A. & Harrison, A. J. M. (1978) Weirs and Flumes for Flow 
Measurement, Wiley. 
Boiten, W. (2000) Hydrometry, Balkema. 
Bos, M. G. (1978) Discharge Measurement Structures, Second Edn, International Institute for Land 
Reclamation and Improvement, Wageningen. 
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, 523–537. 
French, R. H. (1985) Open-Channel Hydraulics, McGraw-Hill, New York. 
Henderson, F. M. (1966) Open Channel Flow, Macmillan, New York. 
Herschy, R. W. (1995) Streamflow Measurement, Second Edn, Spon, London. 
Jaeger, C. (1956) Engineering Fluid Mechanics, Blackie, London. 
Montes, S. (1998) Hydraulics of Open Channel Flow, ASCE, New York. 
2
Open channel hydraulics John Fenton 
Novak, P., Moffat, A. I. B., Nalluri, C. & Narayanan, R. (2001) Hydraulic Structures, Third Edn, Spon, 
London. 
Yalin, M. S. & Ferreira da Silva, A. M. (2001) Fluvial Processes, IAHR, Delft. 
Useful references 
The following table shows some of the many references available, which the lecturer may refer to in 
these notes, or which students might find useful for further reading. For most books in the list, The 
University of Melbourne Engineering Library Reference Numbers are given. 
Reference Comments 
Bos, M. G. (1978), Discharge Measurement Structures, second edn, International Insti-tute 
for Land Reclamation and Improvement, Wageningen. 
Good encyclopaedic treatment 
of structures 
Bos, M. G., Replogle, J. A. & Clemmens, A. J. (1984), Flow Measuring Flumes for 
Open Channel Systems, Wiley. 
Good encyclopaedic treatment 
of structures 
Chanson, H. (1999), The Hydraulics of Open Channel Flow, Arnold, London. Good technical book, moderate 
level, also sediment aspects 
Chaudhry, M. H. (1993), Open-channel flow, Prentice-Hall. Good technical book 
Chow, V. T. (1959), Open-channel Hydraulics, McGraw-Hill, New York. Classic, now dated, not so read-able 
Dooge, J. C. I. (1992) , The Manning formula in context, in B. C. Yen, ed., Channel 
Flow Resistance: Centennial of Manning’s Formula, Water Resources Publications, 
Littleton, Colorado, pp. 136–185. 
Interesting history of Man-ning’s 
law 
Fenton, J. D. & Keller, R. J. (2001), The calculation of streamflow from measurements 
of stage, Technical Report 01/6, Co-operative Research Centre for Catchment Hydrol-ogy, 
Monash University. 
Two level treatment - practical 
aspects plus high level review 
of theory 
Francis, J. & Minton, P. (1984), Civil Engineering Hydraulics, fifth edn, Arnold, Lon-don. 
Good elementary introduction 
French, R. H. (1985), Open-Channel Hydraulics, McGraw-Hill, New York. Wide general treatment 
Henderson, F. M. (1966), Open Channel Flow, Macmillan, New York. Classic, high level, readable 
Hicks, D. M. & Mason, P. D. (1991 ) , Roughness Characteristics of New Zealand 
Rivers, DSIR Marine and Freshwater, Wellington. 
Interesting presentation of 
Manning’s n for different 
streams 
Jain, S. C. (2001), Open-Channel Flow, Wiley. High level, but terse and read-able 
Montes, S. (1998), Hydraulics of Open Channel Flow, ASCE, New York. Encyclopaedic 
Novak, P., Moffat, A. I. B., Nalluri, C. & Narayanan, R. (2001), Hydraulic Structures, 
third edn, Spon, London. 
Standard readable presentation 
of structures 
Townson, J. M. (1991), Free-surface Hydraulics, Unwin Hyman, London. Simple, readable, mathematical 
1. Introduction 
The flow of water with an unconfined free surface at atmospheric pressure presents some of the most 
common problems of fluid mechanics to civil and environmental engineers. Rivers, canals, drainage 
canals, floods, and sewers provide a number of important applications which have led to the theories and 
methods of open channel hydraulics. The main distinguishing characteristic of such studies is that the 
location of the surface is also part of the problem. This allows the existence of waves, both stationary 
and travelling. In most cases, where the waterway is much longer than it is wide or deep, it is possible to 
treat the problem as an essentially one-dimensional one, and a number of simple and powerful methods 
have been developed. 
In this course we attempt a slightly more general view than is customary, where we allow for real fluid 
effects as much as possible by allowing for the variation of velocity over the waterway cross section. We 
recognise that we can treat this approximately, but it remains an often-unknown aspect of each problem. 
3
Open channel hydraulics John Fenton 
This reminds us that we are obtaining approximate solutions to approximate problems, but it does allow 
some simplifications to be made. 
The basic approximation in open channel hydraulics, which is usually a very good one, is that variation 
along the channel is gradual. One of the most important consequences of this is that the pressure in the 
water is given by the hydrostatic approximation, that it is proportional to the depth of water above. 
In Australia there is a slightly non-standard nomenclature which is often used, namely to use the word 
”channel” for a canal, which is a waterway which is usually constructed, and with a uniform section. 
We will use the more international English convention, that such a waterway is called a canal, and we 
will use the words ”waterway”, ”stream”, or ”channel” as generic terms which can describe any type of 
irregular river or regular canal or sewer with a free surface. 
1.1 Types of channel flow to be studied 
(a) Steady uniform flow 
dn 
(b) Steady gradually-varied flow 
dn Normal depth 
(c) Steady rapidly-varied flow 
(d) Unsteady flow 
Figure 1-1. Different types of flow in an open channel 
Case (a) – Steady uniform flow: Steady flow is where there is no change with time, ∂/∂t ≡ 0. 
Distant from control structures, gravity and friction are in balance, and if the cross-section is constant, 
the flow is uniform, ∂/∂x ≡ 0. We will examine empirical laws which predict flow for given bed slope 
and roughness and channel geometry. 
Case (b) – Steady gradually-varied flow: Gravity and friction are in balance here too, but when a 
control is introduced which imposes a water level at a certain point, the height of the surface varies along 
the channel for some distance. For this case we will develop the differential equation which describes 
how conditions vary along the waterway. 
Case (c) – Steady rapidly-varied flow: Figure 1-1(c) shows three separate gradually-varied flow 
states separated by two rapidly-varied regions: (1) flow under a sluice gate and (2) a hydraulic jump. 
The complete problem as presented in the figure is too difficult for us to study, as the basic hydraulic 
approximation that variation is gradual and that the pressure distribution is hydrostatic breaks down in the 
rapid transitions between the different gradually-varied states. We can, however, analyse such problems 
by considering each of the almost-uniform flow states and consider energy or momentum conservation 
between them as appropriate. In these sorts of problems we will assume that the slope of the stream 
4
Open channel hydraulics John Fenton 
balances the friction losses and we treat such problems as frictionless flow over a generally-horizontal 
bed, so that for the individual states between rapidly-varied regions we usually consider the flow to be 
uniform and frictionless, so that the whole problem is modelled as a sequence of quasi-uniform flow 
states. 
Case (d) – Unsteady flow: Here conditions vary with time and position as a wave traverses the 
waterway. We will obtain some results for this problem too. 
1.2 Properties of channel flow 
z = η 
y 
z 
min z = z 
Figure 1-2. Cross-section of flow, showing isovels, contours on which velocity normal to the section is constant. 
Consider a section of a waterway of arbitrary section, as shown in Figure 1-2. The x co-ordinate is 
horizontal along the direction of the waterway (normal to the page), y is transverse, and z is vertical. At 
the section shown the free surface is z = η, which we have shown to be horizontal across the section, 
which is a good approximation in many flows. 
1.2.1 Discharge across a cross-section 
The volume flux or discharge Q at any point is 
Q = 
Z 
A 
u dA = UA 
where u is the velocity component in the x or downstream direction, and A is the cross-sectional area. 
This equation defines the mean horizontal velocity over the section U . In most hydraulic applications 
the discharge is a more important quantity than the velocity, as it is the volume of water and its rate of 
propagation, the discharge, which are important. 
1.2.2 A generalisation – net discharge across a control surface 
Having obtained the expression for volume flux across a plane surface where the velocity vector is 
normal to the surface, we introduce a generalisation to a control volume of arbitrary shape bounded by a 
control surface CS. If u is the velocity vector at any point throughout the control volume and ˆn is a unit 
vector with direction normal to and directed outwards from a point on the control surface, then u · ˆn on 
the control surface is the component of velocity normal to the control surface. If dS is an elemental area 
of the control surface, then the rate at which fluid volume is leaving across the control surface over that 
5
Open channel hydraulics John Fenton 
elemental area is u · ˆndS, and integrating gives 
Total rate at which fluid volume is leaving across the control surface = 
Z 
CS 
u · ˆndS. (1.1) 
If we consider a finite length of channel as shown in Figure 1-3, with the control surface made up of 
u1 
nˆ1 
u2 
nˆ 2 
Figure 1-3. Section of waterway and control surface with vertical ends 
the bed of the channel, two vertical planes across the channel at stations 1 and 2, and an imaginary 
enclosing surface somewhere above the water level, then if the channel bed is impermeable, u · n ˆ= 0 
there; u = 0 on the upper surface; on the left (upstream) vertical plane u · n ˆ= −u1, where u1 is 
the horizontal component of velocity (which varies across the section); and on the right (downstream) 
vertical plane u · n ˆ= +u2. Substituting into equation (1.1) we have 
Z 
Total rate at which fluid volume is leaving across the control surface = − 
A1 
u1 dA + 
Z 
A2 
u2 dA 
= −Q1 + Q2. 
If the flow is steady and there is no increase of volume inside the control surface, then the total rate of 
volume leaving is zero and we have Q1 = Q2. 
While that result is obvious, the results for more general situations are not so obvious, and we will 
generalise this approach to rather more complicated situations – notably where the water surface in the 
Control Surface is changing. 
1.2.3 A further generalisation – transport of other quantities across the control surface 
We saw that u · ˆndS is the volume flux through an elemental area – if wemultiply by fluid density ρ then 
ρ u · ˆndS is the rate at which fluid mass is leaving across an elemental area of the control surface, with 
a corresponding integral over the whole surface. Mass flux is actually more fundamental than volume 
flux, for volume is not necessarily conserved in situations such as compressible flow where the density 
varies. However in most hydraulic engineering applications we can consider volume to be conserved. 
Similarly we can compute the rate at which almost any physical quantity, vector or scalar, is being 
transported across the control surface. For example, multiplying the mass rate of transfer by the fluid 
velocity u gives the rate at which fluid momentum is leaving across the control surface, ρuu · ˆndS. 
1.2.4 The energy equation in integral form for steady flow 
Bernoulli’s theorem states that: 
In steady, frictionless, incompressible flow, the energy per unit mass p/ρ+gz +V 2/2 is constant 
6
Open channel hydraulics John Fenton 
along a streamline, 
where V is the fluid speed, V 2 = u2+v2+w2, inwhich (u, v, w) are velocity components in a cartesian 
co-ordinate system (x, y, z) with z vertically upwards, g is gravitational acceleration, p is pressure and 
ρ is fluid density. In hydraulic engineering it is usually more convenient to divide by g such that we say 
that the head p/ρg + z + V 2/2g is constant along a streamline. 
In open channel flows (and pipes too, actually, but this seems never to be done) we have to consider 
the situation where the energy per unit mass varies across the section (the velocity near pipe walls and 
channel boundaries is smaller than in the middle while pressures and elevations are the same). In this 
case we cannot apply Bernoulli’s theorem across streamlines. Instead, we use an integral form of the 
energy equation, although almost universally textbooks then neglect variation across the flow and refer 
to the governing theorem as ”Bernoulli”. Here we try not to do that. 
The energy equation in integral form can be written for a control volume CV bounded by a control 
surface CS, where there is no heat added or work done on the fluid in the control volume: 
∂ 
∂t 
Z 
CV 
ρ e dV 
| {z } 
Rate at which energy is increasing inside the CV 
+ 
Z 
(p + ρe) u.ˆndS 
| {z } 
CS 
Rate at which energy is leaving the CS 
= 0, (1.2) 
where t is time, ρ is density, dV is an element of volume, e is the internal energy per unit mass of fluid, 
which in hydraulics is the sum of potential and kinetic energies 
e = gz + 
1 
2 
¡ 
u2 + v2 + w2¢ 
, 
where the velocity vector u = (u, v, w) in a cartesian coordinate system (x, y, z) with x horizontally 
along the channel and z upwards, ˆn is a unit vector as above, p is pressure, and dS is an elemental area 
of the control surface. 
Here we consider steady flow so that the first term in equation (1.2) is zero. The equation becomes: 
Z 
CS 
³ 
p + ρgz + 
ρ 
2 
¡ 
u2 + v2 + w2¢´ 
u.ˆn dS = 0. 
We intend to consider problems such as flows in open channels where there is usually no important 
contribution from lateral flows so that we only need to consider flow entering across one transverse face 
of the control surface across a pipe or channel and leaving by another. To do this we have the problem 
of integrating the contribution over a cross-section denoted by A which we also use as the symbol for 
the cross-sectional area. When we evaluate the integral over such a section we will take u to be the 
velocity along the channel, perpendicular to the section, and v and w to be perpendicular to that. The 
contribution over a section of area A is then ±E, where E is the integral over the cross-section: 
E = 
Z 
A 
³ 
p + ρgz + 
ρ 
2 
¡ 
u2 + v2 + w2¢´ 
u dA, (1.3) 
and we take the ± depending on whether the flow is leaving/entering the control surface, because u.ˆn = 
±u. In the case of no losses, E is constant along the channel. The quantity ρQE is the total rate of 
energy transmission across the section. 
Now we consider the individual contributions: 
(a) Velocity head term ρ 
2 
R 
A 
¡ 
u2 + v2 + w2 
¢ 
u dA 
If the flow is swirling, then the v and w components will contribute, and if the flow is turbulent there 
will be extra contributions as well. It seems that the sensible thing to do is to recognise that all velocity 
components and velocity fluctuations will be of a scale given by the mean flow velocity in the stream at 
7
Open channel hydraulics John Fenton 
that point,and so we simply write, for the moment ignoring the coefficient ρ/2: 
Z 
A 
¡ 
u2 + v2 + w2¢ 
u dA = αU 3A = α 
Q3 
A2 , (1.4) 
which defines α as a coefficient which will be somewhat greater than unity, given by 
α = 
R 
A 
¡ 
u2 + v2 + w2 
¢ 
u dA 
U 3A 
. (1.5) 
Conventional presentations define it as being merely due to the non-uniformity of velocity distribution 
across the channel: 
α = 
R 
A u3 dA 
U 3A 
, 
however we suggest that is more properly written containing the other velocity components (and turbu-lent 
contributions as well, ideally). This coefficient is known as a Coriolis coefficient, in honour of the 
French engineer who introduced it. 
Most presentations of open channel theory adopt the approximation that there is no variation of velocity 
over the section, such that it is assumed that α = 1, however that is not accurate. Montes (1998, p27) 
quotes laboratory measurements over a smooth concrete bed giving values of α of 1.035-1.064, while 
for rougher boundaries such as earth channels larger values are found, such as 1.25 for irrigation canals 
in southern Chile and 1.35 in the Rhine River. For compound channels very much larger values may be 
encountered. It would seem desirable to include this parameter in our work, which we will do. 
(b) Pressure and potential head terms 
These are combined as Z 
A 
(p + ρgz) u dA. (1.6) 
The approximation we now make, common throughout almost all open-channel hydraulics, is the ”hy-drostatic 
approximation”, that pressure at a point of elevation z is given by 
p ≈ ρg × height of water above = ρg (η − z) , (1.7) 
where the free surface directly above has elevation η. This is the expression obtained in hydrostatics for 
a fluid which is not moving. It is an excellent approximation in open channel hydraulics except where 
the flow is strongly curved, such as where there are short waves on the flow, or near a structure which 
disturbs the flow. Substituting equation (1.7) into equation (1.6) gives 
ρg 
Z 
A 
η u dA, 
for the combination of the pressure and potential head terms. If we make the reasonable assumption that 
η is constant across the channel the contribution becomes 
ρgη 
Z 
A 
u dA = ρgηQ, 
from the definition of discharge Q. 
(c) Combined terms 
Substituting both that expression and equation (1.4) into (1.3) we obtain 
E = ρgQ 
μ 
η + 
α 
2g 
Q2 
A2 
¶ 
, (1.8) 
8
Open channel hydraulics John Fenton 
which, in the absence of losses, would be constant along a channel. This energy flux across entry and 
exit faces is that which should be calculated, such that it is weighted with respect to the mass flow rate. 
Most presentations pretend that one can just apply Bernoulli’s theorem, which is really only valid along 
a streamline. However our results in the end are not much different. We can introduce the concept of the 
Mean Total Head H such that 
H = 
Energy flux 
g × Mass flux = 
E 
g × ρQ 
= η + 
α 
2g 
Q2 
A2 , (1.9) 
which has units of length and is easily related to elevation in many hydraulic engineering applications, 
relative to an arbitrary datum. The integral version, equation (1.8), is more fundamental, although in 
common applications it is simpler to use the mean total head H, which will simply be referred to as the 
head of the flow. Although almost all presentations of open channel hydraulics assume α = 1, we will 
retain the general value, as a better model of the fundamentals of the problem, which is more accurate, 
but also is a reminder that although we are trying to model reality better, its value is uncertain to a degree, 
and so are any results we obtain. In this way, it is hoped, we will maintain a sceptical attitude to the 
application of theory and ensuing results. 
(d) Application to a single length of channel – including energy losses 
We will represent energy losses by ΔE. For a length of channel where there are no other entry or exit 
points for fluid, we have 
Eout = Ein − ΔE, 
giving, from equation (1.8): 
ρQout 
μ 
gη + 
α 
2 
Q2 
A2 
¶ 
out 
= ρQin 
μ 
gη + 
α 
2 
Q2 
A2 
¶ 
in − ΔE, 
and as there is no mass entering or leaving, Qout = Qin = Q, we can divide through by ρQ and by g, as 
is common in hydraulics: 
μ 
η + 
α 
2g 
Q2 
A2 
¶ 
out 
= 
μ 
η + 
α 
2g 
Q2 
A2 
¶ 
in − ΔH, 
where we have written ΔE = ρgQ × ΔH, where ΔH is the head loss. In spite of our attempts to use 
energy flux, as Q is constant and could be eliminated, in this head form the terms appear as they are used 
in conventional applications appealing to Bernoulli’s theorem, but with the addition of the α coefficients. 
2. Conservation of energy in open channel flow 
In this section and the following one we examine the state of flow in a channel section by calculating the 
energy and momentum flux at that section, while ignoring the fact that the flow at that section might be 
slowly changing. We are essentially assuming that the flow is locally uniform – i.e. it is constant along 
the channel, ∂/∂x ≡ 0. This enables us to solve some problems, at least to a first, approximate, order. 
We can make useful deductions about the behaviour of flows in different sections, and the effects of 
gates, hydraulic jumps, etc.. Often this sort of analysis is applied to parts of a rather more complicated 
flow, such as that shown in Figure 1-1(c) above, where a gate converts a deep slow flow to a faster shallow 
flow but with the same energy flux, and then via an hydraulic jump the flow can increase dramatically in 
depth, losing energy through turbulence but with the same momentum flux. 
2.1 The head/elevation diagram and alternative depths of flow 
Consider a steady (∂/∂t ≡ 0) flow where any disturbances are long, such that the pressure is hydro-static. 
We make a departure from other presentations. Conventionally (beginning with Bakhmeteff in 
1912) they introduce a co-ordinate origin at the bed of the stream and introduce the concept of ”specific 
energy”, which is actually the head relative to that special co-ordinate origin. We believe that the use of 
9
Open channel hydraulics John Fenton 
that datum somehow suggests that the treatment and the results obtained are special in some way. Also, 
for irregular cross-sections such as in rivers, the ”bed” or lowest point of the section is poorly defined, 
and we want to minimise our reliance on such a point. Instead, we will use an arbitrary datum for the 
head, as it is in keeping with other areas of hydraulics and open channel theory. 
Over an arbitrary section such as in Figure 1-2, from equation (1.9), the head relative to the datum can 
be written 
H = η + 
αQ2 
2g 
1 
A2(η) 
, (2.1) 
where we have emphasised that the cross-sectional area for a given section is a known function of surface 
elevation, such that we write A(η). A typical graph showing the dependence of H upon η is shown in 
Figure 2-1, which has been drawn for a particular cross-section and a constant value of discharge Q, 
such that the coefficient αQ2/2g in equation (2.1) is constant. 
η1 
ηc 
η2 
zmin 
Hc 
Surface 
elevation 
η 
H = η + αQ2 
Head H = E/ρgQ 
2g 
1 
A2(η) 
H = η 
1 
2 
Figure 2-1. Variation of head with surface elevation for a particular cross-section and discharge 
The figure has a number of important features, due to the combination of the linear increasing function 
η and the function 1/A2(η) which decreases with η. 
• In the shallow flow limit as η → zmin (i.e. the depth of flow, and hence the cross-sectional area 
A(η), both go to zero while holding discharge constant) the value of H ∼ αQ2/2gA2(η) becomes 
very large, and goes to ∞ in the limit. 
• In the other limit of deep water, as η becomes large, H ∼ η, as the velocity contribution becomes 
negligible. 
• In between these two limits there is a minimum value of head, at which the flow is called critical 
flow, where the surface elevation is ηc and the head Hc. 
• For all other H greater than Hc there are two values of depth possible, i.e. there are two different 
flow states possible for the same head. 
• The state with the larger depth is called tranquil, slow, or sub-critical flow, where the potential to 
make waves is relatively small. 
• The other state, with smaller depth, of course has faster flow velocity, and is called shooting, fast, or 
super-critical flow. There is more wave-making potential here, but it is still theoretically possible 
for the flow to be uniform. 
• The two alternative depths for the same discharge and energy have been called alternate depths. 
10
Open channel hydraulics John Fenton 
That terminology seems to be not quite right – alternate means ”occur or cause to occur by turns, 
go repeatedly from one to another”. Alternative seems better - ”available as another choice”, and 
we will use that. 
• In the vicinity of the critical point, where it is easier for flow to pass from one state to another, the 
flow can very easily form waves (and our hydrostatic approximation would break down). 
• Flows can pass from one state to the other. Consider the flow past a sluice gate in a channel as 
shown in Figure 1-1(c). The relatively deep slow flow passes under the gate, suffering a large 
reduction in momentum due to the force exerted by the gate and emerging as a shallower faster 
flow, but with the same energy. These are, for example, the conditions at the points labelled 1 and 
2 respectively in Figure 2-1. If we have a flow with head corresponding to that at the point 1 with 
surface elevation η1 then the alternative depth is η2 as shown. It seems that it is not possible to 
go in the other direction, from super-critical flow to sub-critical flow without some loss of energy, 
but nevertheless sometimes it is necessary to calculate the corresponding sub-critical depth. The 
mathematical process of solving either problem, equivalent to reading off the depths on the graph, 
is one of solving the equation 
αQ2 
2gA2(η1) 
+ η1 
| {z } 
H1 
= 
αQ2 
2gA2(η2) 
+ η2 
| {z } 
H2 
(2.2) 
for η2 if η1 is given, or vice versa. Even for a rectangular section this equation is a nonlinear tran-scendental 
equation which has to be solved numerically by procedures such as Newton’s method. 
2.2 Critical flow 
δη 
δA 
B 
Figure 2-2. Cross-section of waterway with increment of water level 
We now need to find what the condition for critical flow is, where the head is a minimum. Equation (2.1) 
is 
H = η + 
α 
2g 
Q2 
A2(η) 
, 
and critical flow is when dH/dη = 0: 
dH 
dη 
= 1− 
αQ2 
gA3(η) × 
dA 
dη 
= 0. 
The problem now is to evaluate the derivative dA/dη. From Figure 2-2, in the limit as δη → 0 the 
element of area δA = B δη,such that dA/dη = B, the width of the free surface. Substituting, we have 
the condition for critical flow: 
α 
Q2B 
gA3 = 1. (2.3) 
11
Open channel hydraulics John Fenton 
This can be rewritten as 
α 
(Q/A)2 
g (A/B) 
= 1, 
and as Q/A = U , the mean velocity over the section, and A/B = D, the mean depth of flow, this means 
that 
Critical flow occurs when α 
U 2 
gD 
= 1, that is, when α × 
(Mean velocity)2 
g × Mean depth = 1. (2.4) 
We write this as 
αF 2 = 1 or √αF = 1, (2.5) 
where the symbol F is the Froude number, defined by: 
F = 
Q/A p 
gA/B 
= 
U 
√gD 
= 
Mean velocity 
√g × Mean depth. 
The usual statement in textbooks is that ”critical flow occurs when the Froude number is 1”. We have 
chosen to generalise this slightly by allowing for the coefficient α not necessarily being equal to 1, giving 
αF 2 = 1at critical flow. Any form of the condition, equation (2.3), (2.4) or (2.5) can be used. The mean 
depth at which flow is critical is the ”critical depth”: 
Dc = α 
U 2 
g 
= α 
Q2 
gA2 . (2.6) 
2.3 The Froude number 
The dimensionless Froude number is traditionally used in hydraulic engineering to express the relative 
importance of inertia and gravity forces, and occurs throughout open channel hydraulics. It is relevant 
where the water has a free surface. It almost always appears in the form of αF 2 rather than F . It might 
be helpful here to define F by writing 
F 2 = 
Q2B 
gA3 . 
Consider a calculation where we attempt to quantify the relative importance of kinetic and potential 
energies of a flow – and as the depth is the only vertical scale we have we will use that to express the 
potential energy. We write 
Mean kinetic energy per unit mass 
Mean potential energy per unit mass = 
12 
αU 2 
gD 
2 αF 2, 
= 1 
which indicates something of the nature of the dimensionless number αF 2. 
Flows which are fast and shallow have large Froude numbers, and those which are slow and deep have 
small Froude numbers. For example, consider a river or canal which is 2m deep flowing at 0.5ms−1 
(make some effort to imagine it - we can well believe that it would be able to flow with little surface 
disturbance!). We have 
F = 
U 
√gD ≈ 
0.5 
√10 × 2 
= 0.11 and F 2 = 0.012 , 
and we can imagine that the rough relative importance of the kinetic energy contribution to the potential 
contribution really might be of the order of this 1%. Now consider flow in a street gutter after rain. The 
velocity might also be 0.5ms−1, while the depth might be as little as 2 cm. The Froude number is 
F = 
U 
√gD ≈ 
0.5 
√10 × 0.02 
= 1.1 and F 2 = 1.2 , 
12
Open channel hydraulics John Fenton 
which is just super-critical, and we can easily imagine it to have many waves and disturbances on it due 
to irregularities in the gutter. 
It is clear that αF 2 expresses the scale of the importance of kinetic energy to potential energy, even 
if not in a 1 : 1 manner (the factor of 1/2). It seems that αF 2 is a better expression of the relative 
importance than the traditional use of F . In fact, we suspect that as it always seems to appear in the form 
αF 2 = αU 2/gD, we could define an improved Froude number, Fimproved = αU 2/gD, which explicitly 
recognises (a) that U 2/gD is more fundamental than U/√gD, and (b) that it is the weighted value of u2 
over the whole section, αU 2, which better expresses the importance of dynamic contributions. However, 
we will use the traditional definition F = U/√gD. In tutorials, assignments and exams, unless advised 
otherwise, you may assume α = 1, as has been almost universally done in textbooks and engineering 
practice. However we will retain α as a parameter in these lecture notes, and we recommend it also in 
professional practice. Retaining it will, in general, give more accurate results, but also, retaining it while 
usually not being quite sure of its actual value reminds us that we should not take numerical results as 
accurately or as seriously as we might. Note that, in the spirit of this, we might well use g ≈ 10 in 
practical calculations! 
Rectangular channel 
There are some special simple features of rectangular channels. These are also applicable to wide chan-nels, 
where the section properties do not vary much with depth, and they can be modelled by equivalent 
rectangular channels, or more usually, purely in terms of a unit width. We now find the conditions for 
critical flow in a rectangular section of breadth b and depth h. We have A = bh. From equation (2.3) the 
condition for critical flow for this section is: 
αQ2 
gb2h3 = 1, (2.7) 
but as Q = U bh, this is the condition 
αU 2 
gh 
= 1. (2.8) 
Some useful results follow if we consider the volume flow per unit width q: 
q = 
Q 
b 
= 
U bh 
b 
= U h. (2.9) 
Eliminating Q from (2.7) or U from (2.8) or simply using (2.6) with Dc = hc for the rectangular section 
gives the critical depth, when H is a minimum: 
hc = 
μ 
α 
q2 
g 
¶1/3 
. (2.10) 
This shows that the critical depth hc for rectangular or wide channels depends only on the flow per unit 
width, and not on any other section properties. As for a rectangular channel it is obvious and convenient 
to place the origin on the bed, such that η = h. Then equation (2.1) for critical conditions when H is a 
minimum, H = Hc becomes 
Hc = hc + 
α 
2g 
Q2 
A2c 
= hc + 
α 
2g 
Q2 
b2h2c 
= hc + 
αq2 
2g 
1 
h2c 
, 
and using equation (2.10) to eliminate the q2 term: 
Hc = hc + 
h3c 
2 
1 
h2c 
= 
3 
2 
hc or, hc = 
2 
3 
Hc. (2.11) 
2.4 Water level changes at local transitions in channels 
Now we consider some simple transitions in open channels from one bed condition to another. 
13
Open channel hydraulics John Fenton 
Sub-critical flow over a step in a channel or a narrowing of the channel section: Consider the 
1 2 
Δ 
Figure 2-3. Subcritical flow passing over a rise in the bed 
Surface 
elevation 
η 
¾ 
Head H = E/ρgQ 
Critical constriction 
Constriction 
Upstream section 
1 
2 
4 
3 
2’ 
Figure 2-4. Head/Surface-elevation relationships for three cross-sections 
flow as shown in Figure 2-3. At the upstream section the (H, η) diagram can be drawn as indicated in 
Figure 2-4. Now consider another section at an elevation and possible constriction of the channel. The 
corresponding curve on Figure 2-4 goes to infinity at the higher value of zmin and the curve can be shown 
to be pushed to the right by this raising of the bed and/or a narrowing of the section. At this stage it is not 
obvious that the water surface does drop down as shown in Figure 2-3, but it is immediately explained 
if we consider the point 1 on Figure 2-4 corresponding to the initial conditions. As we assume that no 
energy is lost in travelling over the channel constriction, the surface level must be as shown at point 2 
on Figure 2-4, directly below 1 with the same value of H, and we see how, possibly against expectation, 
the surface really must drop down if subcritical flow passes through a constriction. 
Sub-critical flow over a step or a narrowing of the channel section causing critical flow: Consider 
14
Open channel hydraulics John Fenton 
now the case where the step Δ is high enough and/or the constriction narrow enough that the previously 
sub-critical flow is brought to critical, going from point 1 as before, but this time going to point 2’ on 
Figure 2-4. This shows that for the given discharge, the section cannot be constricted more than this 
amount which would just take it to critical. Otherwise, the (H, η) curve for this section would be moved 
further to the right and there would be no real depth solutions and no flow possible. In this case the 
flow in the constriction would remain critical but the upstream depth would have to increase so as to 
make the flow possible. The step is then acting as a weir, controlling the flow such that there is a unique 
relationship between flow and depth. 
Super-critical flow over a step in a channel or a narrowing of the channel section: Now consider 
super-critical flow over the same constriction as shown in Figure 2-5. In this case the depth actually 
increases as the water passes over the step, going from 3 to 4, as the construction in Figure 2-4 shows. 
Δ 
3 
4 
Figure 2-5. Supercritical flow passing over a hump in the bed. 
Themathematical problem in each of these cases is to solve an equation similar to (2.2) for η2, expressing 
the fact that the head is the same at the two sections: 
αQ2 
2gA21 
(η1) 
+ η1 
| {z } 
H1 
= 
αQ2 
2gA22 
(η2) 
+ η2 
| {z } 
H2 
. (2.12) 
As the relationship between area and elevation at 2 is different from that at 1, we have shown two 
different functions for area as a function of elevation, A1(η1) and A2(η2). 
Example: A rectangular channel of width b1 carries a flow of Q, with a depth h1. The channel 
section is narrowed to a width b2 and the bed raised by Δ, such that the flow depth above the bed 
is now h2. Set up the equation which must be solved for h2. 
Equation (2.12) can be used. If we place the datum on the bed at 1, then η1 = h1 and A1(η1) = 
b1η1 = b1h1. Also, η2 = Δ + h2 and A2(η2) = b2 (η2 − Δ) = b2h2. The equation becomes 
αQ2 
2gb21 
h21 
+ h1 = 
αQ2 
2gb22 
h22 
+ Δ + h2, to be solved for h2, OR, 
αQ2 
2gb21 
h21 
+ h1 = 
αQ2 
2gb22 
(η2 − Δ)2 + η2, to be solved for η2. 
In either case the equation, after multiplying through by h2 or η2 respectively, becomes a cubic, 
which has no simple analytical solution and generally has to be solved numerically. Below we 
will present methods for this. 
2.5 Some practical considerations 
2.5.1 Trapezoidal sections 
15
Open channel hydraulics John Fenton 
γ 
1 
B 
h 
W 
Figure 2-6. Trapezoidal section showing important quantities 
Most canals are excavated to a trapezoidal section, and this is often used as a convenient approximation 
to river cross-sections too. In many of the problems in this course we will consider the case of trapezoidal 
sections. We will introduce the terms defined in Figure 2-6: the bottom width is W , the depth is h, the 
top width is B, and the batter slope, defined to be the ratio of H:V dimensions is γ. From these the 
following important section properties are easily obtained: 
Top width : B = W +2γh 
Area : A = h (W + γh) 
p 
1 + γ2h, 
Wetted perimeter : P = W + 2 
where we will see that the wetted perimeter is an important quantity when we consider friction in chan-nels. 
(Ex. Obtain these relations). 
2.5.2 Solution methods for alternative depths 
Here we consider the problem of solving equation (2.12) numerically: 
αQ2 
2gA21 
(η1) 
+ η1 = 
αQ2 
2gA22 
(η2) 
+ η2, 
where we assume that we know the upstream conditions at point 1 and we have to find η2. The right side 
shows sufficiently complicated dependence on η2 that even for rectangular sections we have to solve this 
problem numerically. Reference can be made to any book on numerical methods for solving nonlinear 
equations, but here we briefly describe some techniques and then develop a simplified version of a robust 
method 
1. Trial and error - evaluate the right side of the equation with various values of η2 until it agreeswith 
the left side. This is simple, but slow to converge and not suitable for machine computation. 
2. Direct iteration - re-arrange the equation in the form 
η2 = H1 − 
αQ2 
2gA22 
(η2) 
and successively evaluate the right side and substitute for η2. We can show that this converges only 
if the flow at 2 is subcritical (αF 2 < 1), the more common case. Provided one is aware of that 
limitation, the method is simple to apply. 
3. Bisection - choose an initial interval in which it is known a solution lies (the value of the function 
changes sign), then successively halve the interval and determine in which half the solution lies each 
time until the interval is small enough. Robust, not quite as simply programmed, but will always 
converge to a solution. 
4. Newton’s method - make an estimate and then make successively better ones by travelling down the 
local tangent. This is fast, and reliable if a solution exists. We write the equation to be solved as 
f (η2) = η2 + 
αQ2 
2gA22 
(η2) − H1 (= 0 when the solution η2 is found). (2.13) 
16
Open channel hydraulics John Fenton 
Then, if η(n) 
2 is the nth estimate of the solution, Newton’s method gives a better estimate: 
η(n+1) 
2 = η(n) 
2 − 
f (η(n) 
2 ) 
f 0(η(n) 
2 ) 
, (2.14) 
where f 0(h2) = ∂f /∂η2. In our case, from (2.13): 
f 0(η2) = 
∂f (η2) 
∂η2 
= 1− 
αQ2 
gA32 
(η2) 
∂A2 
∂η2 
= 1− 
αQ2B2(η2) 
gA(η2) 
32 
= 1− αF 2(η2), 
which is a simple result - obtained using the procedure we used for finding critical flow in an 
arbitrary section. Hence, the procedure (2.14) is 
η(n+1) 
2 = η(n) 
2 − 
η(n) 
2 + αQ2 
2gA22 
2 ) −H1 
(η(n) 
1−αF (n)2 
2 
. (2.15) 
Note that this will not converge as quickly if the flow at 2 is critical, where both numerator and 
denominator go to zero as the solution is approached, but the quotient is still finite. This expression 
looks complicated, but it is simple to implement on a computer, although is too complicated to 
appear on an examination paper in this course. 
These methods will be examined in tutorials. 
2.6 Critical flow as a control - broad-crested weirs 
For a given discharge, the (H, η) diagram showed that the bed cannot be raised or the section narrowed 
more than the amount which would just take it to critical. Otherwise there would be no real depth 
solutions and no flow possible. If the channel were constricted even more, then the depth of flow over 
the raised bed would remain constant at the critical depth, and the upstream depth would have to increase 
so as to make the flow possible. The step is then acting as a weir, controlling the flow. 
hc 
hc 
hc 
Figure 2-7. A broad-crested weir spillway, showing the critical depth over it providing a control. 
Consider the situation shown in Figure 2-7 where the bed falls away after the horizontal section, such as 
on a spillway. The flow upstream is subcritical, but the flow downstream is fast (supercritical). Some-where 
between the two, the flow depth must become critical - the flow reaches its critical depth at some 
point on top of the weir, and the weir provides a control for the flow, such that a relationship between 
flow and depth exists. In this case, the head upstream (the height of the upstream water surface above the 
sill) uniquely determines the discharge, and it is enough to measure the upstream surface elevation where 
the flow is slow and the kinetic part of the head negligible to provide a point on a unique relationship 
between that head over the weir and the discharge. No other surface elevation need be measured. 
Figure 2-8 shows a horizontal flow control, a broad-crested weir, in a channel. In recent years there has 
been a widespread development (but not in Australia, unusually) of such broad-crested weirs placed in 
streams where the flow is subcritical both before and after the weir, but passes through critical on the 
17
Open channel hydraulics John Fenton 
weir. There is a small energy loss after the flume. The advantage is that it is only necessary to measure 
the upstream head over the weir. 
Small energy loss 
hc 
hc hc 
Figure 2-8. A broad-crested weir 
3. Conservation of momentum in open channel flow 
3.1 Integral momentum theorem 
P 
Control volume 
M1 
1 2 
M2 
Figure 3-1. Obstacle in stream reducing the momentum flux 
We have applied energy conservation principles. Now we will apply momentum. We will consider, like 
several problems above, relatively short reaches and channels of prismatic (constant) cross-section such 
that the small contributions due to friction and the component of gravity down the channel are roughly in 
balance. Figure 3-1 shows the important horizontal contributions to force and momentum in the channel, 
where there is a structure applying a force P to the fluid in the control volume we have drawn. 
The momentum theorem applied to the control volume shown can be stated: the net momentum flux 
leaving the control volume is equal to the net force applied to the fluid in the control volume. The 
momentum flux is defined to be the surface integral over the control surface CS: 
Z 
CS 
(p ˆn + u ρu.ˆn) dS, 
where ˆn is a unit vector normal to the surface, such that the pressure contribution on an element of area 
dS is the force p dS times the unit normal vector ˆn giving its direction; u is the velocity vector such 
that u.ˆn is the component of velocity normal to the surface, u.ˆn dS is the volume rate of flow across the 
surface, multiplying by density gives the mass rate of flow across the surface ρu.ˆn dS, and multiplying 
by velocity gives uρu.ˆn dS, the momentum rate of flow across the surface. 
We introduce i, a unit vector in the x direction. On the face 1 of the control surface in Figure 3-1, as the 
outwards normal is in the upstream direction, we have ˆn = −i, and u = u1i, giving u.ˆn = −u1and the 
18
Open channel hydraulics John Fenton 
vector momentum flux across face 1 is 
M1 = −i 
R 
A1 
¡ 
p1 + ρu21 
¢ 
dA = −i M1, 
where the scalar quantity 
M1 = 
R 
A1 
¡ 
p1 + ρu21 
¢ 
dA. 
Similarly, on face 2 of the control surface, as the outwards normal is in the downstream direction, we 
have ˆn = i and u = u2i, giving u.ˆn = +u2 and the vector momentum flux across face 2 is 
M2 = +i 
R 
A2 
¡ 
p2 + ρu22 
¢ 
dA = +i M2 
with scalar quantity 
M2 = 
R 
A2 
¡ 
p2 + ρu22 
¢ 
dA. 
Using the momentum theorem, and recognising that the horizontal component of the force of the body 
on the fluid is −P i, then we have, writing it as a vector equation but including only x (i) components: 
M1 +M2 = −P i 
AsM1 = −M1i andM2 = +M2i, we can write it as a scalar equation giving: 
P = M1 − M2, (3.1) 
where P is the force of the water on the body (or bodies). 
3.1.1 Momentum flux across a section of channel 
From the above, it can be seen how useful is the concept of the horizontal momentum flux at a section of 
the flow in a waterway: 
M = 
Z 
A 
¡ 
p + ρu2¢ 
dA. 
We attach different signs to the contributions depending on whether the fluid is leaving (+ve) or en-tering 
(-ve) the control volume. As elsewhere in these lectures on open channel hydraulics we use the 
hydrostatic approximation for the pressure: p = ρg(η − z), which gives 
M = ρ 
Z 
A 
¡ 
g(η − z) + u2¢ 
dA. 
Now we evaluate this in terms of the quantities at the section. 
Pressure and elevation contribution ρ 
R 
A g(η − z) dA : The integral 
R 
A (η − z) dA is simply the 
first moment of area about a transverse horizontal axis at the surface, we can write it as 
R 
A (η − z) dA = A¯h, (3.2) 
where ¯h 
is the depth of the centroid of the section below the surface. 
Velocity contribution ρ 
R 
A u2 dA : Now we have the task of evaluating the square of the horizontal 
velocity over the section. As with the kinetic energy integral, it seems that the sensible thing to do is 
to recognise that all velocity components and velocity fluctuations will be of a scale given by the mean 
flow velocity in the stream at that point, and so we simply write 
Z 
A 
u2 dA = βU 2A = β 
Q2 
A 
, (3.3) 
19
Open channel hydraulics John Fenton 
which defines β as a coefficient which will be somewhat greater than unity, given by 
β = 
R 
A u2 dA 
U 2A 
. (3.4) 
This coefficient is known as a Boussinesq coefficient, in honour of the French engineer who introduced 
it, who did much important work in the area of the non-uniformity of velocity and the non-hydrostatic 
nature of the pressure distribution. Most presentations of open channel theory adopt the approximation 
that there is no variation of velocity over the section, such that it is assumed that β = 1. Typical real 
values are β = 1.05 − 1.15, somewhat less than the Coriolis energy coefficient α. 
Combining: We can substitute to give the expression we will use for the Momentum Flux: 
M = ρ 
¡ 
gA¯h 
+ βU 2A 
¢ 
= ρ 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
= ρg 
μ 
A(h)¯ h(h) + 
βQ2 
g 
1 
A(h) 
¶ 
(3.5) 
where we have shown the dependence on depth in each term. This expression can be compared with that 
for the head as defined in equation (2.1) but here expressed relative to the bottom of the channel: 
H = h + 
αQ2 
2g 
1 
A2(h) 
. 
The variation with h is different between this and equation (3.5). For large h, H ∼ h, while M ∼ A(h) h(h), which for a rectangular section goes like h2. For small h, H ∼ 1/A2(h), and M ∼ 1/A(h). 
Note that we can re-write equation (3.5) in terms of Froude number (actually appearing as F 2 – yet 
again) to indicate the relative importance of the two parts, which we could think of as ”static” and 
”dynamic” contributions: 
M = ρgA¯h 
μ 
1+βF 2 A/B 
¯h 
¶ 
. 
The ratio (A/B) /¯ h, mean depth to centroid depth, will have a value typically of about 2. 
Example: Calculate (a) Head (using the channel bottom as datum) and (b) Momentum flux, for a 
rectangular section of breadth b and depth h. 
We have A = bh, h = h/2. Substituting into equations (2.1) and (3.5) we obtain 
H = h + 
αQ2 
2gb2 × 
1 
h2 and, 
M = ρ 
μ 
gb 
2 × h2+ 
βQ2 
b × 
1 
h 
¶ 
. 
Note the quite different variation with h between the two quantities. 
3.1.2 Minimum momentum flux and critical depth 
We calculate the condition for minimum M: 
∂M 
∂h 
= 
∂ 
∂h 
(A(h)¯ h(h)) − 
βQ2 
g 
1 
A2(h) 
∂A 
∂h 
= 0. (3.6) 
The derivative of the first moment of area about the surface is obtained by considering the surface 
increased by an amount h + δh 
∂(A¯ h) 
∂h 
= lim 
δh→0 
(A(h)¯ h(h))h+δh − A(h)¯ h(h)) 
δh 
. (3.7) 
The situation is as shown in Figure 3-2. The first moment of area about an axis transverse to the channel 
20
Open channel hydraulics John Fenton 
Depth to centroid of hatched area: δh/2 
B 
h 
Depth to centroid of white area: h +δh 
δh 
Figure 3-2. Geometrical interpretation of calculation of position of centroid 
at the new surface is: 
(A(h)¯ h(h))h+δh = A(h) × (¯h 
+ δh) + B × δh × 
δh 
2 
, 
so that, substituting into equation (3.7), in the limit δh → 0, 
∂(A¯ h) 
∂h 
= lim 
δh→0 
A(h) × (¯h 
+ δh) + B × δh × δh/2 − A(h)¯ h(h)) 
δh 
= A(h) = A, (3.8) 
which is surprisingly simple. Substituting both this and ∂A/∂h = B in equation (3.6), we get the 
condition for minimum M: 
βQ2B 
gA3 = βF 2 = 1, (3.9) 
which is a similar condition for the minimum energy, but as in general α6= β, the condition for minimum 
momentum is not the same as that for minimum energy. 
3.1.3 Momentum flux -depth diagram 
If the cross-section changes or there are other obstacles to the flow, the sides of the channel and/or the 
obstacles will also exert a force along the channel on the fluid. We can solve for the total force exerted 
between two sections if we know the depth at each. In the same way as we could draw an (H, η) diagram 
for a given channel section, we can draw an (M, η) diagram. It is more convenient here to choose the 
datum on the bed of the channel so that we can interpret the surface elevation η as the depth h. Figure 
3-3 shows a momentum flux – depth (M, h) diagram. Note that it shows some of the main features of 
the (H, h) diagram, with two possible depths for the same momentum flux – called conjugate depths. 
However the limiting behaviours for small and large depths are different for momentum, compared with 
energy. 
3.2 Flow under a sluice gate and the hydraulic jump 
Consider the flow problem shown at the top of Figure 3-4, with sub-critical flow (section 0) controlled 
by a sluice gate. The flow emerges from under the gate flowing fast (super-critically, section 1). There 
has been little energy loss in the short interval 0-1, but the force of the gate on the flow has substantially 
reduced its momentum flux. It could remain in this state, however here we suppose that the downstream 
level is high enough such that a hydraulic jump occurs, where there is a violent turbulent motion and in 
a short distance the water changes to sub-critical flow again. In the jump there has been little momentum 
loss, but the turbulence has caused a significant loss of energy between 1-2. After the jump, at stage 2, 
the flow is sub-critical again. We refer to this depth as being sequent to the original depth. 
In the bottom part of Figure 3-4 we combine the (H, h) and (M, h) diagrams, so that the vertical axis is 
21
Open channel hydraulics John Fenton 
Momentum flux M 
Depth h 
h1 
h2 
3 
2 
1 
4 
Rectangular: M ~ h2 
Rectangular: M ~ 1/ h 
P 
h4 P 
h3 
Figure 3-3. Momentum flux – depth diagram, showing effects of a momentum loss P for subcritical and supercrit-ical 
flow. 
0 h 0 
P 
Head H, Momentum flux M 
Depth 
h 
Head H 
Momentum flux M 
0 
2 2 
1 1 
H 0 = H1 M 1 = M 2 M 0 
2 h 
1 h 
P 
0 1 2 
Figure 3-4. Combined Head and Momentum diagrams for the sluice gate and hydraulic jump problem 
22
Open channel hydraulics John Fenton 
depth h and the two horizontal axes are head H and momentum flux M, with different scales. We now 
outline the procedure we follow to analyse the problem of flow under a sluice gate, with upstream force 
P , and a subsequent hydraulic jump. 
• We are given the discharge Q and the upstream depth h0, and we know the cross-sectional details 
of the channel. 
• We can compute the energy and momentum at 0, H0 and M0 (see points 0 on the M −H −h plot). 
• As energy is conserved between 0 and 1, the depth h1 can be calculated by solving the energy 
equation with H1 = H0, possibly using Newton’s method. 
• In fact, this depth may not always be realisable, if the gate is not set at about the right position. The 
flow at the lip of the gate leaves it vertically, and turns around to horizontal, so that the gate opening 
must be larger than h1. A rough guide is that the gate opening must be such that h1 ≈ 0.6×Gate 
opening. 
• With this h1 we can calculate the momentum flux M1. 
• The force on the gate P (assuming that the channel is prismatic) can be calculated from: 
P = M0 − M1 = ρ 
μ 
gAh+β 
Q2 
A 
¶ 
0 − ρ 
μ 
gAh + β 
Q2 
A 
¶ 
1 
• Across the hydraulic jump momentum is conserved, such that M2 = M1: 
μ 
gAh+β 
Q2 
A 
¶ 
2 
= 
μ 
gAh+β 
Q2 
A 
¶ 
1 
• This gives a nonlinear equation for h2 to be solved numerically (note that A and¯h 
are both functions 
of h). In the case of a rectangular channel the equation can be written 
1 
2 
gh21 
+ 
βq2 
h1 
= 
1 
2 
gh22 
+ 
βq2 
h2 
, 
where q = Q/b, the discharge per unit width. In fact it can be solved analytically. Grouping like 
terms on each side and factorising: 
(h2 − h1)(h2 + h1) = 
2βq2 
g 
μ 
1 
h1 − 
1 
h2 
¶ 
, 
h22 
h1 + h21h2 − 
2βq2 
g 
= 0, 
which is a quadratic in h2, with solutions 
h2 = − 
h1 
2 ± 
s 
h21 
4 
+ 
2βq2 
gh1 
, 
but we cannot have a negative depth, and so only the positive sign is taken. Dividing through by 
h1: 
h2 
h1 
= − 
1 
2 
+ 
s 
1 
4 
+ 
2βq2 
gh31 
= 
1 
2 
μq 
1 + 8βF 2 
¶ 
1 −1 
• Sometimes the actual depth of the downstream flow is determined by the boundary condition further 
downstream. If it is not deep enough the actual jump may be an undular hydraulic jump, which 
does not dissipate as much energy, with periodic waves downstream. 
• The pair of depths (h1, h2) for which the flow has the same momentum are traditionally called the 
conjugate depths. 
23
Open channel hydraulics John Fenton 
• The loss in energy H2 − H1 can be calculated. For a rectangular channel it can be shown that 
ΔH = H1 − H2 = 
(h2 − h1)3 
4h1h2 
. 
3.3 The effects of streams on obstacles and obstacles on streams 
3.3.1 Interpretation of the effects of obstacles in a flow 
Slow (sub-critical) approach flow Figure 3-5 shows that the effect of a drag force is to lower the 
P 
P 
h 
1 h 
2 h 
1 2 M 
c h 
Figure 3-5. Effect of obstacles on a subcritical flow 
water surface (counter-intuitive!?) if the flow is slow (sub-critical). 
Fast (super-critical) approach flow Figure 3-6 shows that the effect of a drag force on a super-critical 
P 
h 
2 P h 
1 h 
1 2 M 
c h 
Figure 3-6. Effect of obstacles on a supercritical flow 
flow is to raise the water surface. In fact, the effect of the local force only spreads gradually through 
the stream by turbulent diffusion, and the predicted change in cross-section will apply some distance 
downstream where the flow has become uniform (rather further than in the diagrams here). 
A practical example is the fast flow downstream of a spillway, shown in Figure 3-7, where the flow 
becomes subcritical via a hydraulic jump. If spillway blocks are used, the water level downstream need 
not be as high, possibly with large savings in channel construction. 
3.3.2 Bridge piers - slow approach flow 
Consider flow past bridge piers as shown in Figure 3-8. As the bridge piers extend throughout the flow, 
for the velocity on the pier we will take the mean upstream velocity V = Q/A1, and equation (3.14) can 
24
Open channel hydraulics John Fenton 
P 
2 P h 
2* 
1 
2 
* 
2 h = depth without blocks 
1 h 
1 2 M 
Figure 3-7. Effect of spillway blocks on lowering the water level in a spillway pool 
Plan 
Side elevation 
c h 
1 
2 P 
c h 
1 2 M 
Figure 3-8. Flow past bridge piers and their effect on the flow 
be used. 
3.3.3 Flow in a narrowing channel - choked flow 
We consider cases where the width reduction is more than in a typical bridge pier problem, such that the 
flow in the throat may become critical, the throat becomes a control, and the flow is said to be choked. If 
so, the upstream depth is increased, to produce a larger momentum flux there so that the imposed force 
due to the convergence now just produces critical flow in the throat. In problems such as these, it is very 
helpful to remember that for a rectangular section, equation (2.10): 
hc = 
¡ 
αq2/g 
¢1/3 
, or, re-written, q = 
p 
gh3c 
/α, 
where q = Q/b, the flow per unit width, and also to observe that at critical depth, equation (2.11): 
H = hc + 
αQ2 
2gb2h2c 
= hc + 
αq2 
2gh2c 
= 
3 
2 
hc, so that hc = 
2 
3 
H. 
It is clear that by reducing b, q = Q/b is increased, until in this case, criticality is reached. While 
25
Open channel hydraulics John Fenton 
c h 
Elevation 
c h 
Plan 
Figure 3-9. Flow through contraction sufficiently narrow that it becomes critical 
generally this is not a good thing, as the bridge would then become a control, where there is a relationship 
between flow and depth, this becomes an advantage in flow measurement applications. In critical flow 
flumes only an upstream head is needed to calculate the flow, and the structure is deliberately designed 
to bring about critical depth at the throat. One way of ensuring this is by putting in a rise in the bed at 
the throat. Note that in the diagram the critical depth on the hump is greater than that upstream because 
the width has been narrowed. 
3.3.4 Drag force on an obstacle 
As well as sluices and weirs, many different types of obstacles can be placed in a stream, such as the piers 
of a bridge, blocks on the bed, Iowa vanes, the bars of a trash-rack etc. or possibly more importantly, 
the effects of trees placed in rivers (”Large Woody Debris”), used in their environmental rehabilitation. 
It might be important to know what the forces on the obstacles are, or in flood studies, what effects the 
obstacles have on the river. 
Substituting equation (3.5) into equation (3.1) (P = M1 − M2) gives the expression: 
P = ρ 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
1 − ρ 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
2 
, (3.10) 
so that if we know the depth upstream and downstream of an obstacle, the force on it can be calculated. 
Usually, however, the calculation does not proceed in that direction, as one wants to calculate the effect 
of the obstacle on water levels. The effects of drag can be estimated by knowing the area of the object 
measured transverse to the flow, a, the drag coefficient Cd, and V , themean fluid speed past the object: 
P = 
1 
2 
ρCdV 2a, (3.11) 
and so, substituting into equation (3.10) gives, after dividing by density, 
1 
2 
CdV 2a = 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
1 − 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
2 
. (3.12) 
We will write the velocity V on the obstacle as being proportional to the upstream velocity, such that we 
write 
V 2 = γd 
μ 
Q 
A1 
¶2 
, (3.13) 
26
Open channel hydraulics John Fenton 
where γd is a coefficient which recognises that the velocity which impinges on the object is generally 
not equal to the mean velocity in the flow. For a small object near the bed, γd could be quite small; for 
an object near the surface it will be slightly greater than 1; for objects of a vertical scale that of the whole 
depth, γd ≈ 1. Equation (3.12) becomes 
1 
2 
γd Cd 
Q2 
A21 
a = 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
1 − 
μ 
gA¯h 
+ β 
Q2 
A 
¶ 
2 
(3.14) 
A typical problem is where the downstream water level is given (sub-critical flow, so that the control is 
downstream), and we want to know by how much the water level will be raised upstream if an obstacle 
is installed. As both A1 and h1 are functions of h1, the solution is given by solving this transcendental 
equation for h1. In the spirit of approximation which can be used in open channel hydraulics, and in the 
interest of simplicity and insight, we now obtain an approximate solution. 
3.3.5 An approximate method for estimating the effect of channel obstructions on flooding 
Momentum flux M 
Depth h 
1 h 
3 
2 
1 
4 
4 h P 
3 h 
Tangent to (M,h) curve 
Approximate h1 
Exact h1 
2 h 
Figure 3-10. Momentum flux – depth diagram showing the approximate value of d1 calculated by approximating 
the curve by its tangent at 2. 
Now an approximation to equation (3.14) will be obtained which enables a direct calculation of the 
change in water level due to an obstacle, without solving the transcendental equation. We consider a 
linearised version of the equation, which means that locally we assume a straight-line approximation to 
the momentum diagram, for a small reduction in momentum, as shown in Figure 3-10. 
Consider a small change of surface elevation δh going from section 1 to section 2, and write the expres-sion 
for the downstream area 
A2 = A1 + B1δh. 
It has been shown above (equation 3.8) that 
∂(A¯ h) 
∂h 
= A, 
and so we can write an expression for A2¯h 
2 in terms of A1¯h 
1 and the small change in surface elevation: 
A2¯h 
2 = A1¯h 
1 + δh 
∂(A¯ h) 
∂h 
¯¯¯¯ 
1 
= A1¯h 
1 + δh A1, 
27
Open channel hydraulics John Fenton 
and so equation (3.14) gives us, after dividing through by g: 
1 
2 
γd Cd 
Q2 
gA21 
a = −δh A1 + β 
Q2 
gA1 − β 
Q2 
g (A1 + B1δh) 
= −δh A1 + β 
Q2 
gA1 
à 
1 − 
μ 
1 + 
B1 
A1 
¶ 
−1 
δh 
! 
. 
Now we use a power series expansion in δh to simplify the last term, neglecting terms like (δh)2. For ε 
small, (1 + ε)−1 ≈ 1 − ε, and so 
1 
2 
γd Cd 
Q2 
gA21 
a ≈ −δh A1 + β 
Q2B1 
gA21 
δh. 
We can now solve this to give an explicit approximation for δh: 
δh ≈ 
12 
γd Cd 
Q2 
gA31 
a 
β Q2B1 
gA31 
− 1 
. 
It is simpler to divide both sides by the mean depth A1/B1 to give: 
δh 
A1/B1 
= 
1 
2 γd Cd F 2 
1 
a 
A1 
βF 2 
1 − 1 
. 
We do not have to worry here that for subcritical flow we do not necessarily know the conditions at 
point 1, but instead we know them at the downstream point 2. Within our linearising approximation, we 
can use either the values at 1 or 2 in this expression, and so we generalise by dropping the subscripts 
altogether, so that we write 
δh 
A/B 
= 
1 
2 γd Cd F 2 a 
A 
βF 2 − 1 
= 1 
2 γd Cd 
a 
A × 
F 2 
βF 2 − 1 
. (3.15) 
Thus we see that the relative change of depth (change of depth divided by mean depth) is directly 
proportional to the coefficient of drag and the fractional area of the blockage, as we might expect. The 
result is modified by a term which is a function of the square of the Froude number. For subcritical 
flow the denominator is negative, and so is δh, so that the surface drops, as we expect, and as can be 
seen when we solve the problem exactly using the momentum diagram. If upstream is supercritical, 
the surface rises. Clearly, if the flow is near critical (βF 2 
1 ≈ 1) the change in depth will be large (the 
gradient on the momentum diagram is vertical), when the theory will have limited validity. 
Example: In a proposal for the rehabilitation of a river it is proposed to install a number of logs 
(”Large Woody Debris” or ”Engineered Log Jam”). If a single log of diameter 500mm and 10m 
long were placed transverse to the flow, calculate the effect on river height. The stream is roughly 
100m wide, say 10m deep in a severe flood, with a drag coefficient Cd ≈ 1. The all-important 
velocities are a bit uncertain. We might assume a mean velocity of say 6ms−1, and velocity on 
the log of 2ms−1. Assume β = 1.1. 
We have the values 
A = 100 × 10 = 1000m2, a = 0.5 × 10 = 5m2 
F 2 = U 2/gD = 62/10/10 = 0.36, γd = 22/62 ≈ 0.1 
and substituting into equation (3.15) gives 
δh 
A/B ≈ 
1 
2 γdCd 
a 
A1 
β − 1 
F 2 
1 
= 
1 
2 × 0.1 × 1 × 5 
1000 
1.1 − 1 
0.36 
= −1.5 × 10−4, 
so that multiplying by the mean depth, δη = −1.5 × 10−4 × 10 = −1.5mm. The negative value is the 
28
Open channel hydraulics John Fenton 
change as we go downstream, thus we see that the flow upstream is raised by 1.5mm. 
4. Uniform flow in prismatic channels 
Uniform flow is where the depth does not change along the waterway. For this to occur the channel 
properties also must not change along the stream, such that the channel is prismatic, and this occurs only 
in constructed canals. However in rivers if we need to calculate a flow or depth, it is common to use 
a cross-section which is representative of the reach being considered, and to assume it constant for the 
application of this theory. 
4.1 Features of uniform flow and relationships for uniform flow 
• There are two forces in balance in steady flow: 
– The component of gravity downstream along the channel, and 
– the shear stress at the sides which offers resistance to the flow, which increases with flow veloc-ity. 
• If a channel is long and prismatic (slope and section do not change) then far from the effects of 
controls the two can be in balance, and if the flow is steady, the mean flow velocity and flow depth 
remain constant along the channel, giving uniform flow, at normal depth. 
A 
L 
P 
τ0 
τ0 
θ 
g sin θ 
g 
Figure 4-1. Slice of uniform channel flow showing shear forces and body forces per unit mass acting 
Consider a slice of uniform flow in a channel of length L and cross-sectional area A, as shown in 
Figure 4-1. The component of gravity force along the channel is ρ × AL × g sin θ, where θ is the 
angle of inclination of the channel, assumed positive downwards. The shear force is τ 0 × L × P , 
where τ 0 is the shear stress, and P is the wetted perimeter of the cross-section. As the two are in 
balance for uniform flow, we obtain 
τ 0 
ρ 
= g 
A 
P 
sin θ. 
Now, τ 0/ρ has units of velocity squared; we combine g and the coefficient relating the mean 
29
Open channel hydraulics John Fenton 
velocity U at a section to that velocity, giving Chézy’s law (1768): 
U = C 
p 
RS0, 
where C is the Chézy coefficient (with units L1/2T−1), R = A/P is the hydraulic radius (L), and 
S0 = sinθ is the slope of the bed, positive downwards. The tradition in engineering is that we use 
the tangent of the slope angle, so this is valid for small slopes such that sin θ ≈ tan θ. 
• However there is experimental evidence that C depends on the hydraulic radius in the form C ∼ R1/6 (Gauckler, Manning), and the law widely used is Manning’s Law: 
U = 
1 
n 
R2/3S1/2 
0 , 
where n is the Manning coefficient (units of L−1/3T), which increases with increasing roughness. 
Typical values are: concrete - 0.013, irrigation channels - 0.025, clean natural streams - 0.03, 
streams with large boulders - 0.05, streams with many trees - 0.07. Usually the units are not shown. 
• Multiplying by the area, Manning’s formula gives the discharge: 
Q = U A = 
1 
n 
A5/3 
P 2/3 
p 
S0, (4.1) 
in which both A and P are functions of the flow depth. Similarly, Chézy’s law gives 
Q = C 
A3/2 
P 1/2 
p 
S0. (4.2) 
Both equations show how flow increases with cross-sectional area and slope and decreases with 
wetted perimeter. 
4.2 Computation of normal depth 
If the discharge, slope, and the appropriate roughness coefficient are known, either of equations (4.1) 
and (4.2) is a transcendental equation for the normal depth hn, which can be solved by the methods 
described earlier. We can gain some insight and develop a simple scheme by considering a trapezoidal 
cross-section, where the bottom width is W , the depth is h, and the batter slopes are (H:V) γ : 1 (see 
Figure 2-6). The following properties are easily shown to hold (the results have already been presented 
above): 
Top width B W +2γh 
Area A h(W + γh) 
Wetted perimeter P W +2 
p 
1 + γ2h 
In the case of wide channels, (i.e. channels rather wider than they are deep, h ¿ W , which is a common 
case) the wetted perimeter does not show a lot of variation with depth h. Similarly in the expression for 
the area, the second factor W +γh (the mean width) does not show a lot of variation with h either – most 
of the variation is in the first part h. Hence, if we assume that these properties hold for cross-sections of 
a more general nature, we can rewrite Manning’s law: 
Q = 
1 
n 
A5/3(h) 
P 2/3(h) 
p 
S0 = 
√S0 
n 
(A(h)/h)5/3 
P 2/3(h) × h5/3, 
where most of the variation with h is contained in the last term h5/3, and by solving for that term we can 
re-write the equation in a form suitable for direct iteration 
h = 
μ 
Qn 
√S0 
¶3/5 
× 
P 2/5(h) 
A(h)/h 
, 
30
Open channel hydraulics John Fenton 
where the first term on the right is a constant for any particular problem, and the second term is expected 
to be a relatively slowly-varying function of depth, so that the whole right side varies slowly with depth – 
a primary requirement that the direct iteration scheme be convergent and indeed be quickly convergent. 
Experience with typical trapezoidal sections shows that this works well and is quickly convergent. How-ever, 
it also works well for flow in circular sections such as sewers, where over a wide range of depths 
the mean width does not vary much with depth either. For small flows and depths in sewers this is not 
so, and a more complicated method might have to be used. 
Example: Calculate the normal depth in a trapezoidal channel of slope 0.001, Manning’s coef-ficient 
n = 0.04, width 10m, with batter slopes 2 : 1, carrying a flow of 20m3 s−1. We have 
A = h (10 + 2 h), P = 10+4.472 h, giving the scheme 
h = 
μ 
Qn 
√S0 
¶3/5 
× 
(10 + 4.472 h)2/5 
10 + 2 h 
= 6.948 × 
(10 + 4.472 h)2/5 
10 + 2 h 
and starting with h = 2 we have the sequence of approximations: 2.000, 1.609, 1.639, 1.637 - 
quite satisfactory in its simplicity and speed. 
4.3 Conveyance 
It is often convenient to use the conveyance K which contains all the roughness and cross-section prop-erties, 
such that for steady uniform flow 
Q = K 
p 
S0, 
such that, using an electrical analogy, the flow (current) is given by a ”conductance” (here conveyance) 
multiplied by a driving potential, which, here in this nonlinear case, is the square root of the bed slope. 
In more general non-uniform flows below we will see that we use the square root of the head gradient. 
With this definition, if we use Manning’s law for the flow, K is defined by 
K = 
1 
n × A 
μ 
A 
P 
¶2/3 
= 
1 
n × 
A5/3 
P 2/3 
, (4.3) 
where K is a function of the roughness and the local depth and cross-section properties. Textbooks often 
use conveyance to provide methods for computing the equivalent conveyance of compound sections such 
as that shown in Figure 4-2. However, for such cases where a river has overflowed its banks, the flow 
situation is much more likely to be more two-dimensional than one-dimensional. The extent of the 
various elemental areas and the Manning’s roughnesses of the different parts are all such as to often 
render a detailed ”rational” calculation unjustified. 
2 3 
1 
Figure 4-2. 
In the compound channel in the figure, even though the surface might actually be curved as shown and 
31
Open channel hydraulics John Fenton 
the downstream slope and/or bed slope might be different across the channel, the tradition is that we 
assume it to be the same. The velocities in the individual sections are, in general, different. We write 
Manning’s law for each section based on the mean bed slope: 
Q1 = K1S1/2 
0 , Q2 = K2S1/2 
0 , Q3 = K3S1/2 
0 
In a general case with n sub-sections, the total discharge is 
Q = 
Xn 
i=1 
Qi = 
Xn 
i=1 
KiS1/2 
0 = S1/2 
0 
Xn 
i=1 
Ki = S1/2 
0 K 
where we use the symbol K for the total conveyance: 
K = 
Xn 
i=1 
Ki = 
Xn 
i=1 
A5/3 
i 
niP 2/3 
i 
. 
5. Steady gradually-varied non-uniform flow 
Steady gradually-varied flow is where the conditions (possibly the cross-section, but often just the sur-face 
elevation) vary slowly along the channel but do not change with time. The most common situation 
where this arises is in the vicinity of a control in a channel, where there may be a structure such as a 
weir, which has a particular discharge relationship between the water surface level and the discharge. 
Far away from the control, the flow may be uniform, and there the relationship between surface elevation 
and discharge is in general a different one, typically being given by Manning’s law, (4.1). The transition 
between conditions at the control and where there is uniform flow is described by the gradually-varied 
flow equation, which is an ordinary differential equation for the water surface height. The solution will 
approach uniform flow if the channel is prismatic, but in general we can treat non-prismatic waterways 
also. 
In sub-critical flow the flow is relatively slow, and the effects of any control can propagate back up the 
channel, and so it is that the numerical solution of the gradually-varied flow equation also proceeds in 
that direction. On the other hand, in super-critical flow, all disturbances are swept downstream, so that 
the effects of a control cannot be felt upstream, and numerical solution also proceeds downstream from 
the control. 
Solution of the gradually-varied flow equation is a commonly-encountered problem in open channel 
hydraulics, as it is used to determine, for example, how far upstream water levels might be increased, 
and hence flooding enhanced, due to downstream works, such as the installation of a bridge. 
5.1 Derivation of the gradually-varied flow equation 
Consider the elemental section of waterway of length Δx shown in Figure 5-1. We have shown stations 
1 and 2 in what might be considered the reverse order, but we will see that for the more common sub-critical 
flow, numerical solution of the governing equation will proceed back up the stream. Considering 
stations 1 and 2: 
Total head at 2 = H2 
Total head at 1 = H1 = H2 − HL, 
and we introduce the concept of the friction slope Sf which is the gradient of the total energy line such 
that HL = Sf × Δx. This gives 
H1 = H2 − Sf Δx, 
32
Open channel hydraulics John Fenton 
U 2 / 2 g Total energy line 
α 2 
U 2 / 2 g 
α 1 
2 h 
S f Δx 
S0 Δx 
2 1 
1 h 
Sub-critical flow 
Δx 
Figure 5-1. Elemental section of waterway 
and if we introduce the Taylor series expansion for H1: 
H1 = H2 + Δx 
dH 
dx 
+ . . . , 
substituting and taking the limit Δx → 0 gives 
dH 
dx 
= −Sf , (5.1) 
an ordinary differential equation for the head as a function of x. 
To obtain the frictional slope, we use either of the frictional laws of Chézy orManning (or a smooth-wall 
formula), where we make the assumption that the equation may be extended from uniform flow (where 
the friction slope equals the constant bed slope) to this non-uniform case, such that the discharge at any 
point is given by, for the case of Manning: 
Q = 
1 
n 
A5/3 
P 2/3 
p 
Sf , 
but where we have used the friction slope Sf rather than bed slope S0, as in uniform flow. Solving for 
Sf : the friction slope is given by 
Sf = 
Q2 
K2(h) 
, (5.2) 
where we have used the conveyance K, which was defined in equation (4.3), but we repeat here, 
K (h) = 
1 
n 
A5/3 
P 2/3 
, 
showing the section properties to be a function of the local depth, where we have restricted our attention 
to prismatic channels on constant slope. This now means that for a given constant discharge we can 
write the differential equation (5.1) as 
dH 
dx 
= −Sf (h). (5.3) 
As we have had to use local depth on the right side, we have to show the head to be a function of depth 
h, so that we write 
H = h + zmin + 
α 
2g 
Q2 
A2(h) 
. (5.4) 
33
Open channel hydraulics John Fenton 
Differentiating: 
dH 
dx 
= 
dh 
dx 
+ 
dzmin 
dx − 
α 
g 
Q2 
A3(h) 
dA(h) 
dx 
. (5.5) 
The derivative dzmin/dx = −S0, where S0 is the bed slope,which we have defined to be positive for the 
usual case of a downwards-sloping channel. Now we have to express the dA(h)/dx in terms of other 
quantities. In our earlier work we saw that if the surface changed by an amount Δh, then the change in 
area due to this was ΔA = B Δh, and so we can write dA(h)/dx = B dh/dx, and substituting these 
results into equation (5.5) gives 
dH 
dx 
= −S0 + 
μ 
1 − 
α 
g 
Q2B(h) 
A3(h) 
¶ 
dh 
dx 
= −S0 + 
¡ 
1 − αF 2(h) 
¢ dh 
dx 
, 
where the Froude number has entered, shown here as a function of depth. Finally, substituting into (5.3) 
we obtain 
dh 
dx 
= 
S0 − Sf (h) 
1 − αF 2(h) 
= 
S0 − Q2/K2(h) 
1 − αF 2(h) 
, (5.6) 
a differential equation for depth h as a function of x, where on the right we have shown the functional 
dependence of the various terms. This, or the less-explicit form (5.3), are forms of the gradually-varied 
flow equation, from which a number of properties can be inferred. 
5.2 Properties of gradually-varied flow and the governing equation 
• The equation and its solutions are important, in that they tell us how far the effects of a structure or 
works in or on a stream extend upstream or downstream. 
• It is an ordinary differential equation of first order, hence one boundary condition must be supplied 
to obtain the solution. In sub-critical flow, this is the depth at a downstream control; in super-critical 
flow it is the depth at an upstream control. 
• In general that boundary depth is not equal to the normal depth, and the differential equation de-scribes 
the transition from the boundary depth to normal depth – upstream for sub-critical flow, 
downstream for supercritical flow. The solutions look like exponential decay curves, and below we 
will show that they are, to a first approximation. 
• If that approximation is made, the resulting analytical solution is useful in providing us with some 
insight into the quantities which govern the extent of the upstream or downstream influence. 
• The differential equation is nonlinear, and the dependence on h is complicated, such that analytical 
solution is not possible without an approximation, and we will usually use numerical methods. 
• The uniform flow limit satisfies the differential equation, for when Sf = S0, dh/dx = 0, and the 
depth does not change. 
• As the flow approaches critical flow, when αF 2 → 1, then dh/dx → ∞, and the surface becomes 
vertical. This violates the assumption we made that the flow is gradually varied and the pressure 
distribution is hydrostatic. This is the one great failure of our open channel hydraulics at this level, 
that it cannot describe the transition between sub- and super-critical flow. 
5.3 Classification system for gradually-varied flows 
The differential equation can be used as the basis for a dual classification system of gradually-varied 
flows: 
• one based on 5 conditions for slope, essentially as to how the normal depth compares with critical 
depth, and 3 conditions for the actual depth, and how it compares with both normal, and critical 
depths, as shown in the Table: 
34
Open channel hydraulics John Fenton 
Slope classification 
Steep slope: hn < hc 
Critical slope: hn = hc 
Mild slope: hn > hc 
Horizontal slope: hn = ∞ 
Adverse slope: hn does not exist 
Depth classification 
Zone 1: h > hn and hc 
Zone 2: h between hn and hc 
Zone 3: h < hn and hc 
Figure 5-2 shows the behaviour of the various solutions. In practice, the most commonly encountered 
are the M1, the backwater curve on a mild slope; M2, the drop-down curve on a mild slope, and S2, the 
drop-down curve on a steep slope. 
5.4 Some practical considerations 
5.4.1 Flood inundation studies 
Figure 5-3 shows a typical subdivision of a river and its flood plain for a flood inundation study, where 
solution of the gradually-varied flow equation would be required. It might be wondered how the present 
methods can be used for problems which are unsteady, such as the passage of a substantial flood, where 
on the front face of the flood wave the water surface is steeper and on the back face it is less steep. In 
many situations, however, the variation of the water slope about the steady slope is relatively small, and 
the wavelength of the flood is long, so that the steady model can be used as a convenient approxima-tion. 
The inaccuracies of knowledge of the geometry and roughness are probably such as to mask the 
numerical inaccuracies of the solution. Below we will present some possible methods and compare their 
accuracy. 
5.4.2 Incorporation of losses 
It is possible to incorporate the losses due, say, to a sudden expansion or contraction of the channel, such 
as shown in Figure 5-4. After an expansion the excess velocity head is destroyed through turbulence. 
Before an expansion the losses will not be so large, but there will be some extra losses due to the 
convergence and enhanced friction. We assume that the expansion/contraction head loss can be written 
ΔHe = C 
μ 
Q2 
2gA22 
− 
Q2 
2gA21 
¶ 
, 
where C ≈ 0.3 for expansions and 0.1 for a contraction. 
5.5 Numerical solution of the gradually-varied flow equation 
Consider the gradually-varied flow equation (5.6) 
dh 
dx 
= 
S0 − Sf (h) 
1 − αF 2(h) 
, 
where both Sf (h) = Q2/K2(h) and F 2(h) = Q2B(h)/gA3(h) are functions of Q as well as the depth 
h. However as Q is constant for a particular problem we do not show the functional dependence on 
it. The equation is a differential equation of first order, and to obtain solutions it is necessary to have a 
boundary condition h = h0 at a certain x = x0, which will be provided by a control. The problem may 
be solved using any of a number of methods available for solving ordinary differential equations which 
35
Open channel hydraulics John Fenton 
Figure 5-2. Typical gradually-varied flow surface profiles, drawn by Dr I. C. O’Neill. 
36
Open channel hydraulics John Fenton 
Typical cross-section used for 1-D analysis 
Edge of flood plain / Extent of 1% flood 
River banks 
Figure 5-3. Practical river problem with subdivision 
2 1 
Figure 5-4. Flow separation and head loss due to a contraction 
are described in books on numerical methods. These methods are usually accurate and can be found 
in many standard software packages. It is surprising that books on open channels do not recognise that 
the problem of numerical solution of the gradually-varied flow equation is actually a standard numerical 
problem, although practical details may make it more complicated. Instead, such texts use methods 
such as the ”Direct step method” and the ”Standard step method”. There are several software packages 
such as HEC-RAS which use such methods, but solution of the gradually-varied flow equation is not a 
difficult problem to solve for specific problems in practice if one knows that it is merely the solution of 
a differential equation, and here we briefly set out the nature of such schemes. 
The direction of solution is very important. If the different conventional cases in Figure 5-2 are exam-ined, 
it can be seen that for the mild slope (sub-critical flow) cases that the surface decays somewhat 
exponentially to normal depth upstream from a downstream control, whereas for steep slope (super-critical 
flow) cases the surface decays exponentially to normal depth downstream from an upstream 
control. This means that to obtain numerical solutions we will always solve (a) for sub-critical flow: 
from the control upstream, and (b) for super-critical flow: from the control downstream. 
5.5.1 Euler’s method 
The simplest (Euler) scheme to advance the solution from (xi, hi) to (xi + Δxi, hi+1) is 
xi+1 ≈ xi + Δxi, where Δxi is negative for subcritical flow, (5.7) 
dh 
hi+1 ≈ hi + Δxi 
dx 
¯¯¯¯ 
i 
= hi + Δxi 
S0 − Sf (hi) 
1 − αF 2(hi) 
. (5.8) 
This is the simplest but least accurate of all methods – yet it might be appropriate for open channel 
37
Open channel hydraulics John Fenton 
problems where quantities may only be known approximately. One can use simple modifications such 
as Heun’s method to gain better accuracy – or even more simply, just take smaller steps Δxi. 
5.5.2 Heun’s method 
In this case the value of hi+1 calculated from equation (5.8) is used as a first estimate h∗i+1, then the 
value of the right hand side of the differential equation is also calculated there, and the mean of the two 
values taken. That is, 
xi+1 ≈ xi + Δxi, again where Δxi is negative for subcritical flow, (5.9) 
h∗i+1 = hi + Δxi 
S0 − Sf (hi) 
1 − αF 2(hi) 
, (5.10) 
hi+1 = hi + 
Δxi 
2 
μ 
S0 − Sf (hi) 
1 − αF 2(hi) 
+ 
S0 − Sf (h∗i+1) 
1 − αF 2(h∗i+1) 
¶ 
. (5.11) 
Neither of these two methods are presented in hydraulics textbooks as alternatives. Although they are 
simple and flexible, they are not as accurate as other less-convenient methods described further below. 
The step Δxi can be varied at will, to suit possible irregularly spaced cross-sectional data. 
5.5.3 Predictor-corrector method – Trapezoidal method 
This is simply an iteration of the last method, whereby the step in equation (5.11) is repeated several 
times, at each stage setting h∗i+1 equal to the updated value of hi+1. This gives an accurate and conve-nient 
method, and it is surprising that it has not been used. 
5.5.4 Direct step method 
Textbooks do present the Direct Step method, which is applied by taking steps in the height and calcu-lating 
the corresponding step in x. It is only applicable to problems where the channel is prismatic. The 
reciprocal of equation (5.1) is 
dx 
dH 
= − 
1 
Sf 
, 
which is then approximated by a version of Heun’s method, but which is not a correct rational approxi-mation: 
Δx = − 
ΔH 
¯ Sf 
, (5.12) 
where a mean value of the friction slope is used. The procedure is: for the control point x0 and h0, 
calculate H0 from equation (5.4), then assume a finite value of depth change Δh to compute h1 = 
h0 + Δh, from which H1 is calculated from equation (5.4), giving ΔH = H1 − H0. Then with ¯ Sf = 
(Sf (h0)+Sf (h1))/2, equation (5.12) is used to calculate the corresponding Δx, giving x1 = x0 +Δx. 
The process is then repeated to give x2 and h2 and so on. It is important to choose the correct sign 
of Δh such that computations proceed in the right direction such that, for example, Δx is negative for 
sub-critical flow, and computations proceed upstream. 
The method has the theoretical disadvantage that it is an inconsistent approximation, in that it should 
actually be computing the mean of 1/Sf, namely 1/Sf , rather than 1/ ¯ Sf . More importantly it has 
practical disadvantages, such that it is applicable only to prismatic sections, results are not obtained at 
specified points in x, and as uniform flow is approached the Δx become infinitely large. However it is a 
surprisingly accurate method. 
5.5.5 Standard step method 
This is an implicit method, requiring numerical solution of a transcendental equation at each step. It 
can be used for irregular channels, and is rather more general. In this case, the distance interval Δx is 
specified and the corresponding depth change calculated. In the Standard step method the procedure is 
38
Open channel hydraulics John Fenton 
to write 
ΔH = −Sf Δx, 
and then write it as 
H2(h2) − H1(h1) = − 
Δx 
2 
(Sf1 + Sf2) , 
for sections 1 and 2, where the mean value of the friction slope is used. This gives 
α 
Q2 
2gA22 
+ Z2 + h2 = α 
Q2 
2gA21 
+ Z1 + h1 − 
Δx 
2 
(Sf1 + Sf2) , 
where Z1 and Z2 are the elevations of the bed. This is a transcendental equation for h2, as this determines 
A2, P2, and Sf2. Solution could be by any of the methods we have had for solving transcendental 
equations, such as direct iteration, bisection, or Newton’s method. 
Although the Standard step method is an accurate and stable approximation, the lecturer considers it 
unnecessarily complicated, as it requires solution of a transcendental equation at each step. It would be 
much simpler to use a simple explicit Euler or Heun’s method as described above. 
Example: Consider a simple backwater problem to test the accuracy of the various methods. A 
trapezoidal channel with bottom width W = 10m, side batter slopes of 2:1, is laid on a slope of 
S0 = 10−4, and carries a flow of Q = 15m3 
s−1. Manning’s coefficient is n = 0.025. At the 
downstream control the depth is 2.5m. Calculate the surface profile (and how far the effect of the 
control extends upstream). Use 10 computational steps over a length of 30km. 
2.5 
2.4 
2.3 
2.2 
2.1 
2 
-45 -40 -35 -30 -25 -20 -15 -10 -5 0 
Depth (m) 
x (km) 
Accurate solution 
Analytical approximation 
Trapezoidal 
Standard step 
Direct step 
Figure 5-5. Comparison of different solution methods – depth plotted. 
Figure 5-5 shows the results of the computations, where depth is plotted, while Figure 5-6 shows the 
same results, but where surface elevation is plotted, to show what the surface profile actually looks 
like. For relatively few computational points Euler’s method was not accurate, and neither was Heun’s 
method, and have not been plotted. The basis of accuracy is shown by the solid line, from a highly-accurate 
Runge-Kutta 4th order method. This is not recommended as a method, however, as it makes use 
of information from three intermediate points at each step, information which in non-prismatic channels 
is not available. It can be seen that the relatively simple Trapezoidal method is sufficiently accurate, 
certainly of acceptable practical accuracy. The Direct Step method was slightly more accurate, but the 
results show one of its disadvantages, that the distance between computational points becomes large as 
uniform flow is approached, and the points are at awkward distances. The last plotted point is at about 
−25km; using points closer to normal depth gave inaccurate results. The Standard Step method was 
very accurate, but is not plotted as it is complicated to apply. Of course, if more computational points 
39
Open channel hydraulics John Fenton 
6.5 
6 
5.5 
5 
4.5 
4 
3.5 
3 
2.5 
2 
Accurate solution and normal depth 
Analytical approximation 
-45 -40 -35 -30 -25 -20 -15 -10 -5 0 
Surface elevation (m) 
x (km) 
Trapezoidal 
Standard step 
Direct step 
Figure 5-6. Comparison of different solution methods – elevation plotted. 
were taken, more accurate results could be obtained. In this example we deliberately chose relatively 
few steps (10) so that the numerical accuracies of the methods could be compared. 
Also plotted on the figures is a dotted line corresponding to the analytical solution which will be de-veloped 
below. Although this was not as accurate as the numerical solutions, it does give a simple 
approximate result for the rate of decay and how far upstream the effects of the control extend. For 
many practical problems, this accuracy and simplicity may be enough. 
The channel dimensions are typical of a large irrigation canal in the Murray Valley - it is interesting that 
the effects of the control extend for some 30km! 
To conclude with a recommendation: the trapezoidal method, Heun’s method iterated several times 
is simple, accurate, and convenient. If, however, a simple approximate solution is enough, then the 
following analytical solution can be used. 
5.6 Analytical solution 
Whereas the numerical solutions give us numbers to analyse, sometimes very few actual numbers are 
required, such as merely requiring how far upstream water levels are raised to a certain level, the effect 
of downstream works on flooding, for example. Here we introduce a different way of looking at a 
physical problem in hydraulics, where we obtain an approximate mathematical solution so that we can 
provide equations which reveal to us more of the nature of the problem than do numbers. Sometimes an 
understanding of what is important is more useful than numbers. 
Consider the water surface depth to be written 
h(x) = h0 + h1(x), 
where we use the symbol h0 for the constant normal depth, and h1(x) is a relatively small departure 
of the surface from the uniform normal depth. We use the governing differential equation (5.6) but we 
assume that the Froude number squared is sufficiently small that it can be ignored. This is not essential, 
but it makes the equations simpler to write and read. (As an example, consider a typical stream flowing 
at 0.5 m/s with a depth of 2m, giving F 2 = 0.0125 - there are many cases where F 2 can be neglected). 
40
Open channel hydraulics John Fenton 
The simplified differential equation can be written 
dh 
dx 
= S0 − S(h), 
where for purposes of simplicity we have dropped the subscript f on the friction slope, now represented 
by S. Substituting our expansion, we obtain 
dh1 
dx 
= S0 − S(h0 + h1(x)). (5.13) 
Now we introduce the approximation that the h1 term is relatively small such that we can write for the 
friction term its Taylor expansion about normal flow: 
S(h0 + h1(x)) = S(h0) + h1(x) × 
dS 
dh 
(h0) + Terms proportional to h21 
. 
We ignore the quadratic terms, write dS/dh(h0) as S/ 
0 , and substituting into equation (5.13), we obtain 
dh1 
dx 
= −S/ 
0 h1 
where we have used S(h0) = S0. This is an ordinary differential equation which we can solve analyt-ically. 
We have achieved this by ”linearising” about the uniform flow. Now, by separation of variables 
we can obtain the solution 
h1 = Ge−S/ 
0 x, 
and the full solution is 
h = h0 + Ge−S/ 
0 x, (5.14) 
where G is a constant which would be evaluated by satisfying the boundary condition at the control. 
This shows that the water surface is actually approximated by an exponential curve passing from the 
value of depth at the control to normal depth. In fact, we will see that as S/ 
0 is negative, far upstream as 
x → −∞, 
the water surface approaches normal depth. 
Now we obtain an expression for S/ 
0 in terms of the channel dimensions. From Manning’s law, 
S = n2Q2 P 4/3 
A10/3 , 
and differentiating gives 
S/ = n2Q2 
à 
4 
3 
P 1/3 
A10/3 
dP 
dh − 
10 
3 
P 4/3 
A13/3 
dA 
dh 
! 
, 
which we can factorise, substitute dA/dh = B, and recognising the term outside the brackets, we obtain 
an analytical expression for the coefficient of x in the exponential function: 
0 = n2Q2 P 4/3 
−S/ 
0 
A10/3 
0 
μ 
10 
3 
B 
A − 
4 
3 
dP/dh 
P 
¶¯¯¯¯ 
0 
= S0 
μ 
10 
3 
B0 
A0 − 
4 
3 
dP0/dh0 
P0 
¶ 
. 
The larger this number, the more rapid is the decay with x. The formula shows that more rapid decay 
occurs with steeper slopes (large S0), smaller depths (B0/A0 = 1/D0, where D0 is the mean depth - 
if it decreases the overall coefficient increases), and smaller widths (P0 is closely related to width, the 
term involving it can be written d(log P0)/dh0: if P0 decreases the term decreases - relatively slowly - 
but the negative sign means that the effect is to increase the magnitude of the overall coefficient). Hence, 
generally the water surface approaches normal depth more quickly for steeper, shallower and narrower 
(i.e. steeper and smaller) streams. The free surface will decay to 10% of its original departure from 
41
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels
Open channels

More Related Content

What's hot

Chapter 4 seepage theories
Chapter 4 seepage theoriesChapter 4 seepage theories
Chapter 4 seepage theories
Mohsin Siddique
 
Permiability
PermiabilityPermiability
Permiability
NaveenHanchinal
 
Cross drainage work
Cross drainage workCross drainage work
Cross drainage work
Divya Vishnoi
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signed
Irfan Malik
 
Soil slope stability
Soil slope stabilitySoil slope stability
(Part i)- open channels
(Part i)- open channels(Part i)- open channels
(Part i)- open channels
Mohsin Siddique
 
River engineering
River engineeringRiver engineering
River engineering
Wollega University
 
Notches and weir
Notches and weirNotches and weir
Notches and weir
Dewang Agrawal
 
Darcys law
Darcys lawDarcys law
Darcys law
Imran Rakib
 
Prefabricated vertical drains
Prefabricated vertical drainsPrefabricated vertical drains
Prefabricated vertical drains
MadhuReddy163523
 
Flow through Mouthpieces
Flow through MouthpiecesFlow through Mouthpieces
Flow through Mouthpieces
Dheeraj Kumar Soni
 
Spillways
SpillwaysSpillways
Cross drainage work
Cross drainage work Cross drainage work
Cross drainage work
meetmaniya
 
Spillways
SpillwaysSpillways
Spillways
Waleed Liaqat
 
Dam outlet works lecture 3
Dam outlet works lecture 3Dam outlet works lecture 3
Dam outlet works lecture 3
Vidhi Khokhani
 
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdfSILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
R K DHIMAN
 
Orifice and mouthpieces
Orifice and mouthpieces Orifice and mouthpieces
Orifice and mouthpieces
Thim Mengly(ម៉េងលី,孟李)
 
Bridge Lecture Slide by Micotol
Bridge Lecture Slide by MicotolBridge Lecture Slide by Micotol
Bridge Lecture Slide by Micotol
Micotol
 
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
abdqadeer
 
Class 5 Permeability Test ( Geotechnical Engineering )
Class 5   Permeability Test ( Geotechnical Engineering )Class 5   Permeability Test ( Geotechnical Engineering )
Class 5 Permeability Test ( Geotechnical Engineering )
Hossam Shafiq I
 

What's hot (20)

Chapter 4 seepage theories
Chapter 4 seepage theoriesChapter 4 seepage theories
Chapter 4 seepage theories
 
Permiability
PermiabilityPermiability
Permiability
 
Cross drainage work
Cross drainage workCross drainage work
Cross drainage work
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signed
 
Soil slope stability
Soil slope stabilitySoil slope stability
Soil slope stability
 
(Part i)- open channels
(Part i)- open channels(Part i)- open channels
(Part i)- open channels
 
River engineering
River engineeringRiver engineering
River engineering
 
Notches and weir
Notches and weirNotches and weir
Notches and weir
 
Darcys law
Darcys lawDarcys law
Darcys law
 
Prefabricated vertical drains
Prefabricated vertical drainsPrefabricated vertical drains
Prefabricated vertical drains
 
Flow through Mouthpieces
Flow through MouthpiecesFlow through Mouthpieces
Flow through Mouthpieces
 
Spillways
SpillwaysSpillways
Spillways
 
Cross drainage work
Cross drainage work Cross drainage work
Cross drainage work
 
Spillways
SpillwaysSpillways
Spillways
 
Dam outlet works lecture 3
Dam outlet works lecture 3Dam outlet works lecture 3
Dam outlet works lecture 3
 
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdfSILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
SILT FACTOR FOR SCOUR CALCULATION AROUND BRIDGE FOUNDATION.pdf
 
Orifice and mouthpieces
Orifice and mouthpieces Orifice and mouthpieces
Orifice and mouthpieces
 
Bridge Lecture Slide by Micotol
Bridge Lecture Slide by MicotolBridge Lecture Slide by Micotol
Bridge Lecture Slide by Micotol
 
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
• Sediment Transport • The Earthen Canal Section • Lined Canals & Its Advantages
 
Class 5 Permeability Test ( Geotechnical Engineering )
Class 5   Permeability Test ( Geotechnical Engineering )Class 5   Permeability Test ( Geotechnical Engineering )
Class 5 Permeability Test ( Geotechnical Engineering )
 

Viewers also liked

Open Channel Flows (Lecture notes 04)
Open Channel Flows (Lecture notes 04)Open Channel Flows (Lecture notes 04)
Open Channel Flows (Lecture notes 04)
Shekh Muhsen Uddin Ahmed
 
(Part iii)- open channels
(Part iii)- open channels(Part iii)- open channels
(Part iii)- open channels
Mohsin Siddique
 
open channel flow
open channel flowopen channel flow
open channel flow
GeoRuizO
 
Open channelhydraulics2
Open channelhydraulics2Open channelhydraulics2
Open channelhydraulics2
Chaminda Rathnayaka
 
Chap1 open channel flow
Chap1 open channel flowChap1 open channel flow
Chap1 open channel flow
UTM International Campus
 
Cc303 hydraulics 1_29.5
Cc303 hydraulics 1_29.5Cc303 hydraulics 1_29.5
Cc303 hydraulics 1_29.5
mumyzz
 
Water level sensors repot
Water level sensors repotWater level sensors repot
Water level sensors repot
Sai Bhaskar Reddy Nakka
 
Coeficiente de coriolis
Coeficiente de coriolisCoeficiente de coriolis
Coeficiente de coriolis
Brayan Jimenez Herrera
 
High powered messaging with RabbitMQ
High powered messaging with RabbitMQHigh powered messaging with RabbitMQ
High powered messaging with RabbitMQ
James Carr
 
Gradually-Varied Flow in Open Channels ( Lecture notes 06)
Gradually-Varied Flow in Open Channels ( Lecture notes 06)Gradually-Varied Flow in Open Channels ( Lecture notes 06)
Gradually-Varied Flow in Open Channels ( Lecture notes 06)
Shekh Muhsen Uddin Ahmed
 
Open channel design
Open channel designOpen channel design
Open channel design
Mohsin Siddique
 
Dymanics of open channel flow
Dymanics of open channel flowDymanics of open channel flow
Dymanics of open channel flow
Mohsin Siddique
 
Ce533 ch1
Ce533 ch1Ce533 ch1
Ce533 ch1
sam4111
 

Viewers also liked (13)

Open Channel Flows (Lecture notes 04)
Open Channel Flows (Lecture notes 04)Open Channel Flows (Lecture notes 04)
Open Channel Flows (Lecture notes 04)
 
(Part iii)- open channels
(Part iii)- open channels(Part iii)- open channels
(Part iii)- open channels
 
open channel flow
open channel flowopen channel flow
open channel flow
 
Open channelhydraulics2
Open channelhydraulics2Open channelhydraulics2
Open channelhydraulics2
 
Chap1 open channel flow
Chap1 open channel flowChap1 open channel flow
Chap1 open channel flow
 
Cc303 hydraulics 1_29.5
Cc303 hydraulics 1_29.5Cc303 hydraulics 1_29.5
Cc303 hydraulics 1_29.5
 
Water level sensors repot
Water level sensors repotWater level sensors repot
Water level sensors repot
 
Coeficiente de coriolis
Coeficiente de coriolisCoeficiente de coriolis
Coeficiente de coriolis
 
High powered messaging with RabbitMQ
High powered messaging with RabbitMQHigh powered messaging with RabbitMQ
High powered messaging with RabbitMQ
 
Gradually-Varied Flow in Open Channels ( Lecture notes 06)
Gradually-Varied Flow in Open Channels ( Lecture notes 06)Gradually-Varied Flow in Open Channels ( Lecture notes 06)
Gradually-Varied Flow in Open Channels ( Lecture notes 06)
 
Open channel design
Open channel designOpen channel design
Open channel design
 
Dymanics of open channel flow
Dymanics of open channel flowDymanics of open channel flow
Dymanics of open channel flow
 
Ce533 ch1
Ce533 ch1Ce533 ch1
Ce533 ch1
 

Similar to Open channels

Chapter 1
Chapter 1Chapter 1
Chapter 1
nimcan1
 
Fluid Mechanics - Fluid Dynamics
Fluid Mechanics - Fluid DynamicsFluid Mechanics - Fluid Dynamics
Fluid Mechanics - Fluid Dynamics
Malla Reddy University
 
lm201-Project
lm201-Projectlm201-Project
lm201-Project
Laura Merchant
 
Materi Oprasi Waduk " Developmen in Water Science"ts "
Materi Oprasi Waduk " Developmen in Water Science"ts "Materi Oprasi Waduk " Developmen in Water Science"ts "
Materi Oprasi Waduk " Developmen in Water Science"ts "
ssuser68f408
 
CHAPTER NO 2
CHAPTER NO 2 CHAPTER NO 2
CHAPTER NO 2
Hadiqa Qadir
 
Chapter 1 introduction to hydraulics structures history...
Chapter  1 introduction to hydraulics structures history...Chapter  1 introduction to hydraulics structures history...
Chapter 1 introduction to hydraulics structures history...
Mohsin Siddique
 
1202.6542
1202.65421202.6542
1202.6542
AemiroGedefaw1
 
Rating curve design,practice and problems
Rating curve design,practice and problemsRating curve design,practice and problems
Rating curve design,practice and problems
Tianjin University China
 
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptxUNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
SumanManandhar9
 
Hydraulics & hydraulic machinery
Hydraulics & hydraulic machineryHydraulics & hydraulic machinery
Hydraulics & hydraulic machinery
Selva Prakash
 
NCIT civil Syllabus 2013-2014
NCIT civil Syllabus 2013-2014NCIT civil Syllabus 2013-2014
NCIT civil Syllabus 2013-2014
LAKSHITHA RAJAPAKSHA
 
Testo
TestoTesto
Lecture 1.pptx
Lecture 1.pptxLecture 1.pptx
Lecture 1.pptx
Robert Zedd
 
S01.s1 material (1)
S01.s1 material (1)S01.s1 material (1)
S01.s1 material (1)
Sergio Ramos
 
OPEN CHANNELS
OPEN CHANNELS OPEN CHANNELS
OPEN CHANNELS
KhurshidAlam124
 
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
Todd Lewis
 
Open Channel Flow
Open Channel FlowOpen Channel Flow
Open Channel Flow
Malla Reddy University
 
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavitLattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
IAEME Publication
 
THIẾT KẾ ĐƯỜNG ỐNG
THIẾT KẾ ĐƯỜNG ỐNGTHIẾT KẾ ĐƯỜNG ỐNG
THIẾT KẾ ĐƯỜNG ỐNG
hieudk53
 
CFD Analysis Of Multi-Phase Flow And Its Measurements
CFD Analysis Of Multi-Phase Flow And Its MeasurementsCFD Analysis Of Multi-Phase Flow And Its Measurements
CFD Analysis Of Multi-Phase Flow And Its Measurements
IOSR Journals
 

Similar to Open channels (20)

Chapter 1
Chapter 1Chapter 1
Chapter 1
 
Fluid Mechanics - Fluid Dynamics
Fluid Mechanics - Fluid DynamicsFluid Mechanics - Fluid Dynamics
Fluid Mechanics - Fluid Dynamics
 
lm201-Project
lm201-Projectlm201-Project
lm201-Project
 
Materi Oprasi Waduk " Developmen in Water Science"ts "
Materi Oprasi Waduk " Developmen in Water Science"ts "Materi Oprasi Waduk " Developmen in Water Science"ts "
Materi Oprasi Waduk " Developmen in Water Science"ts "
 
CHAPTER NO 2
CHAPTER NO 2 CHAPTER NO 2
CHAPTER NO 2
 
Chapter 1 introduction to hydraulics structures history...
Chapter  1 introduction to hydraulics structures history...Chapter  1 introduction to hydraulics structures history...
Chapter 1 introduction to hydraulics structures history...
 
1202.6542
1202.65421202.6542
1202.6542
 
Rating curve design,practice and problems
Rating curve design,practice and problemsRating curve design,practice and problems
Rating curve design,practice and problems
 
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptxUNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
UNIT 7_The Groundwater System_Kathmandu University_DrSuman2022.pptx
 
Hydraulics & hydraulic machinery
Hydraulics & hydraulic machineryHydraulics & hydraulic machinery
Hydraulics & hydraulic machinery
 
NCIT civil Syllabus 2013-2014
NCIT civil Syllabus 2013-2014NCIT civil Syllabus 2013-2014
NCIT civil Syllabus 2013-2014
 
Testo
TestoTesto
Testo
 
Lecture 1.pptx
Lecture 1.pptxLecture 1.pptx
Lecture 1.pptx
 
S01.s1 material (1)
S01.s1 material (1)S01.s1 material (1)
S01.s1 material (1)
 
OPEN CHANNELS
OPEN CHANNELS OPEN CHANNELS
OPEN CHANNELS
 
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
Hydraulic Design of Energy Dissipators for Culverts and Channels, 3rd Edition...
 
Open Channel Flow
Open Channel FlowOpen Channel Flow
Open Channel Flow
 
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavitLattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
Lattice boltzmann simulation of non newtonian fluid flow in a lid driven cavit
 
THIẾT KẾ ĐƯỜNG ỐNG
THIẾT KẾ ĐƯỜNG ỐNGTHIẾT KẾ ĐƯỜNG ỐNG
THIẾT KẾ ĐƯỜNG ỐNG
 
CFD Analysis Of Multi-Phase Flow And Its Measurements
CFD Analysis Of Multi-Phase Flow And Its MeasurementsCFD Analysis Of Multi-Phase Flow And Its Measurements
CFD Analysis Of Multi-Phase Flow And Its Measurements
 

Recently uploaded

🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
dulbh kashyap
 
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
shourabjaat424
 
SPICE PARK JUL2024 ( 6,866 SPICE Models )
SPICE PARK JUL2024 ( 6,866 SPICE Models )SPICE PARK JUL2024 ( 6,866 SPICE Models )
SPICE PARK JUL2024 ( 6,866 SPICE Models )
Tsuyoshi Horigome
 
SELENIUM CONF -PALLAVI SHARMA - 2024.pdf
SELENIUM CONF -PALLAVI SHARMA - 2024.pdfSELENIUM CONF -PALLAVI SHARMA - 2024.pdf
SELENIUM CONF -PALLAVI SHARMA - 2024.pdf
Pallavi Sharma
 
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
Geoffrey Wardle. MSc. MSc. Snr.MAIAA
 
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
sexytaniya455
 
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
DharmaBanothu
 
Data Communication and Computer Networks Management System Project Report.pdf
Data Communication and Computer Networks Management System Project Report.pdfData Communication and Computer Networks Management System Project Report.pdf
Data Communication and Computer Networks Management System Project Report.pdf
Kamal Acharya
 
Sri Guru Hargobind Ji - Bandi Chor Guru.pdf
Sri Guru Hargobind Ji - Bandi Chor Guru.pdfSri Guru Hargobind Ji - Bandi Chor Guru.pdf
Sri Guru Hargobind Ji - Bandi Chor Guru.pdf
Balvir Singh
 
Better Builder Magazine, Issue 49 / Spring 2024
Better Builder Magazine, Issue 49 / Spring 2024Better Builder Magazine, Issue 49 / Spring 2024
Better Builder Magazine, Issue 49 / Spring 2024
Better Builder Magazine
 
anatomy of space vehicle and aerospace structures.pptx
anatomy of space vehicle and aerospace structures.pptxanatomy of space vehicle and aerospace structures.pptx
anatomy of space vehicle and aerospace structures.pptx
ssusercf1619
 
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
hotchicksescort
 
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort ServiceCuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
yakranividhrini
 
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
Tsuyoshi Horigome
 
Basic principle and types Static Relays ppt
Basic principle and  types  Static Relays pptBasic principle and  types  Static Relays ppt
Basic principle and types Static Relays ppt
Sri Ramakrishna Institute of Technology
 
Call Girls Madurai 8824825030 Escort In Madurai service 24X7
Call Girls Madurai 8824825030 Escort In Madurai service 24X7Call Girls Madurai 8824825030 Escort In Madurai service 24X7
Call Girls Madurai 8824825030 Escort In Madurai service 24X7
Poonam Singh
 
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
sonamrawat5631
 
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
AK47
 
paper relate Chozhavendhan et al. 2020.pdf
paper relate Chozhavendhan et al. 2020.pdfpaper relate Chozhavendhan et al. 2020.pdf
paper relate Chozhavendhan et al. 2020.pdf
ShurooqTaib
 
Covid Management System Project Report.pdf
Covid Management System Project Report.pdfCovid Management System Project Report.pdf
Covid Management System Project Report.pdf
Kamal Acharya
 

Recently uploaded (20)

🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
🚺ANJALI MEHTA High Profile Call Girls Ahmedabad 💯Call Us 🔝 9352988975 🔝💃Top C...
 
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
Call Girls Chandigarh 🔥 7014168258 🔥 Real Fun With Sexual Girl Available 24/7...
 
SPICE PARK JUL2024 ( 6,866 SPICE Models )
SPICE PARK JUL2024 ( 6,866 SPICE Models )SPICE PARK JUL2024 ( 6,866 SPICE Models )
SPICE PARK JUL2024 ( 6,866 SPICE Models )
 
SELENIUM CONF -PALLAVI SHARMA - 2024.pdf
SELENIUM CONF -PALLAVI SHARMA - 2024.pdfSELENIUM CONF -PALLAVI SHARMA - 2024.pdf
SELENIUM CONF -PALLAVI SHARMA - 2024.pdf
 
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
My Aerospace Design and Structures Career Engineering LinkedIn version Presen...
 
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
Call Girls Nagpur 8824825030 Escort In Nagpur service 24X7
 
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
An In-Depth Exploration of Natural Language Processing: Evolution, Applicatio...
 
Data Communication and Computer Networks Management System Project Report.pdf
Data Communication and Computer Networks Management System Project Report.pdfData Communication and Computer Networks Management System Project Report.pdf
Data Communication and Computer Networks Management System Project Report.pdf
 
Sri Guru Hargobind Ji - Bandi Chor Guru.pdf
Sri Guru Hargobind Ji - Bandi Chor Guru.pdfSri Guru Hargobind Ji - Bandi Chor Guru.pdf
Sri Guru Hargobind Ji - Bandi Chor Guru.pdf
 
Better Builder Magazine, Issue 49 / Spring 2024
Better Builder Magazine, Issue 49 / Spring 2024Better Builder Magazine, Issue 49 / Spring 2024
Better Builder Magazine, Issue 49 / Spring 2024
 
anatomy of space vehicle and aerospace structures.pptx
anatomy of space vehicle and aerospace structures.pptxanatomy of space vehicle and aerospace structures.pptx
anatomy of space vehicle and aerospace structures.pptx
 
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
❣Unsatisfied Bhabhi Call Girls Surat 💯Call Us 🔝 7014168258 🔝💃Independent Sura...
 
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort ServiceCuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
Cuttack Call Girls 💯Call Us 🔝 7374876321 🔝 💃 Independent Female Escort Service
 
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
Update 40 models( Solar Cell ) in SPICE PARK(JUL2024)
 
Basic principle and types Static Relays ppt
Basic principle and  types  Static Relays pptBasic principle and  types  Static Relays ppt
Basic principle and types Static Relays ppt
 
Call Girls Madurai 8824825030 Escort In Madurai service 24X7
Call Girls Madurai 8824825030 Escort In Madurai service 24X7Call Girls Madurai 8824825030 Escort In Madurai service 24X7
Call Girls Madurai 8824825030 Escort In Madurai service 24X7
 
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
🔥Young College Call Girls Chandigarh 💯Call Us 🔝 7737669865 🔝💃Independent Chan...
 
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
Call Girls In Rohini (Delhi) Call 9711199012 ∰ Escort Service In Delhi ∰
 
paper relate Chozhavendhan et al. 2020.pdf
paper relate Chozhavendhan et al. 2020.pdfpaper relate Chozhavendhan et al. 2020.pdf
paper relate Chozhavendhan et al. 2020.pdf
 
Covid Management System Project Report.pdf
Covid Management System Project Report.pdfCovid Management System Project Report.pdf
Covid Management System Project Report.pdf
 

Open channels

  • 1. April 18, 2007 Open channel hydraulics John Fenton Abstract This course of 15 lectures provides an introduction to open channel hydraulics, the generic name for the study of flows in rivers, canals, and sewers, where the distinguishing characteristic is that the surface is unconfined. This means that the location of the surface is also part of the problem, and allows for the existence of waves – generally making things more interesting! At the conclusion of this subject students will understand the nature of flows and waves in open channels and be capable of solving a wide range of commonly encountered problems. Table of Contents References . . . . . . . . . . . . . . . . . . . . . . . 2 1. Introduction . . . . . . . . . . . . . . . . . . . . . 3 1.1 Types of channel flowtobestudied . . . . . . . . . . . . 4 1.2 Properties of channel flow . . . . . . . . . . . . . . 5 2. Conservation of energy in open channel flow . . . . . . . . . . . 9 2.1 The head/elevation diagram and alternative depths of flow . . . . . 9 2.2 Critical flow . . . . . . . . . . . . . . . . . . . 11 2.3 TheFroudenumber . . . . . . . . . . . . . . . . 12 2.4 Waterlevelchangesatlocaltransitionsinchannels . . . . . . . 13 2.5 Somepracticalconsiderations . . . . . . . . . . . . . 15 2.6 Critical flowasacontrol-broad-crestedweirs . . . . . . . . 17 3. Conservation of momentum in open channel flow . . . . . . . . . 18 3.1 Integralmomentumtheorem . . . . . . . . . . . . . . 18 3.2 Flow under a sluice gate and the hydraulic jump . . . . . . . . 21 3.3 The effects of streams on obstacles and obstacles on streams . . . . 24 4. Uniform flowinprismaticchannels . . . . . . . . . . . . . . 29 4.1 Features of uniform flow and relationships for uniform flow . . . . 29 4.2 Computationofnormaldepth . . . . . . . . . . . . . 30 4.3 Conveyance . . . . . . . . . . . . . . . . . . . 31 5. Steady gradually-varied non-uniform flow . . . . . . . . . . . 32 5.1 Derivation of the gradually-varied flowequation . . . . . . . . 32 5.2 Properties of gradually-varied flow and the governing equation . . . 34 5.3 Classification system for gradually-varied flows . . . . . . . . 34 1
  • 2. Open channel hydraulics John Fenton 5.4 Somepracticalconsiderations . . . . . . . . . . . . . 35 5.5 Numerical solution of the gradually-varied flowequation . . . . . 35 5.6 Analyticalsolution . . . . . . . . . . . . . . . . . 40 6. Unsteady flow . . . . . . . . . . . . . . . . . . . . 42 6.1 Massconservationequation . . . . . . . . . . . . . . 42 6.2 Momentum conservation equation – the low inertia approximation . . 43 6.3 Diffusion routing and nature of wave propagation in waterways . . . 45 7. Structures in open channels and flowmeasurement . . . . . . . . . 47 7.1 Overshotgate-thesharp-crestedweir . . . . . . . . . . . 47 7.2 Triangularweir . . . . . . . . . . . . . . . . . . 48 7.3 Broad-crested weirs – critical flowasacontrol . . . . . . . . 48 7.4 Freeoverfall . . . . . . . . . . . . . . . . . . . 49 7.5 Undershotsluicegate . . . . . . . . . . . . . . . . 49 7.6 Drownedundershotgate . . . . . . . . . . . . . . . 50 7.7 DethridgeMeter . . . . . . . . . . . . . . . . . 50 8. The measurement of flowinriversandcanals . . . . . . . . . . 50 8.1 Methodswhichdonotusestructures . . . . . . . . . . . 50 8.2 The hydraulics of a gauging station . . . . . . . . . . . . 53 8.3 Ratingcurves . . . . . . . . . . . . . . . . . . 54 9. Loose-boundary hydraulics . . . . . . . . . . . . . . . . 56 9.1 Sedimenttransport . . . . . . . . . . . . . . . . . 56 9.2 Incipientmotion . . . . . . . . . . . . . . . . . 57 9.3 Turbulent flowinstreams . . . . . . . . . . . . . . . 58 9.4 Dimensionalsimilitude . . . . . . . . . . . . . . . 58 9.5 Bed-loadrateoftransport–Bagnold’sformula . . . . . . . . 59 9.6 Bedforms . . . . . . . . . . . . . . . . . . . 59 References Ackers, P., White, W. R., Perkins, J. A. & Harrison, A. J. M. (1978) Weirs and Flumes for Flow Measurement, Wiley. Boiten, W. (2000) Hydrometry, Balkema. Bos, M. G. (1978) Discharge Measurement Structures, Second Edn, International Institute for Land Reclamation and Improvement, Wageningen. 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, 523–537. French, R. H. (1985) Open-Channel Hydraulics, McGraw-Hill, New York. Henderson, F. M. (1966) Open Channel Flow, Macmillan, New York. Herschy, R. W. (1995) Streamflow Measurement, Second Edn, Spon, London. Jaeger, C. (1956) Engineering Fluid Mechanics, Blackie, London. Montes, S. (1998) Hydraulics of Open Channel Flow, ASCE, New York. 2
  • 3. Open channel hydraulics John Fenton Novak, P., Moffat, A. I. B., Nalluri, C. & Narayanan, R. (2001) Hydraulic Structures, Third Edn, Spon, London. Yalin, M. S. & Ferreira da Silva, A. M. (2001) Fluvial Processes, IAHR, Delft. Useful references The following table shows some of the many references available, which the lecturer may refer to in these notes, or which students might find useful for further reading. For most books in the list, The University of Melbourne Engineering Library Reference Numbers are given. Reference Comments Bos, M. G. (1978), Discharge Measurement Structures, second edn, International Insti-tute for Land Reclamation and Improvement, Wageningen. Good encyclopaedic treatment of structures Bos, M. G., Replogle, J. A. & Clemmens, A. J. (1984), Flow Measuring Flumes for Open Channel Systems, Wiley. Good encyclopaedic treatment of structures Chanson, H. (1999), The Hydraulics of Open Channel Flow, Arnold, London. Good technical book, moderate level, also sediment aspects Chaudhry, M. H. (1993), Open-channel flow, Prentice-Hall. Good technical book Chow, V. T. (1959), Open-channel Hydraulics, McGraw-Hill, New York. Classic, now dated, not so read-able Dooge, J. C. I. (1992) , The Manning formula in context, in B. C. Yen, ed., Channel Flow Resistance: Centennial of Manning’s Formula, Water Resources Publications, Littleton, Colorado, pp. 136–185. Interesting history of Man-ning’s law Fenton, J. D. & Keller, R. J. (2001), The calculation of streamflow from measurements of stage, Technical Report 01/6, Co-operative Research Centre for Catchment Hydrol-ogy, Monash University. Two level treatment - practical aspects plus high level review of theory Francis, J. & Minton, P. (1984), Civil Engineering Hydraulics, fifth edn, Arnold, Lon-don. Good elementary introduction French, R. H. (1985), Open-Channel Hydraulics, McGraw-Hill, New York. Wide general treatment Henderson, F. M. (1966), Open Channel Flow, Macmillan, New York. Classic, high level, readable Hicks, D. M. & Mason, P. D. (1991 ) , Roughness Characteristics of New Zealand Rivers, DSIR Marine and Freshwater, Wellington. Interesting presentation of Manning’s n for different streams Jain, S. C. (2001), Open-Channel Flow, Wiley. High level, but terse and read-able Montes, S. (1998), Hydraulics of Open Channel Flow, ASCE, New York. Encyclopaedic Novak, P., Moffat, A. I. B., Nalluri, C. & Narayanan, R. (2001), Hydraulic Structures, third edn, Spon, London. Standard readable presentation of structures Townson, J. M. (1991), Free-surface Hydraulics, Unwin Hyman, London. Simple, readable, mathematical 1. Introduction The flow of water with an unconfined free surface at atmospheric pressure presents some of the most common problems of fluid mechanics to civil and environmental engineers. Rivers, canals, drainage canals, floods, and sewers provide a number of important applications which have led to the theories and methods of open channel hydraulics. The main distinguishing characteristic of such studies is that the location of the surface is also part of the problem. This allows the existence of waves, both stationary and travelling. In most cases, where the waterway is much longer than it is wide or deep, it is possible to treat the problem as an essentially one-dimensional one, and a number of simple and powerful methods have been developed. In this course we attempt a slightly more general view than is customary, where we allow for real fluid effects as much as possible by allowing for the variation of velocity over the waterway cross section. We recognise that we can treat this approximately, but it remains an often-unknown aspect of each problem. 3
  • 4. Open channel hydraulics John Fenton This reminds us that we are obtaining approximate solutions to approximate problems, but it does allow some simplifications to be made. The basic approximation in open channel hydraulics, which is usually a very good one, is that variation along the channel is gradual. One of the most important consequences of this is that the pressure in the water is given by the hydrostatic approximation, that it is proportional to the depth of water above. In Australia there is a slightly non-standard nomenclature which is often used, namely to use the word ”channel” for a canal, which is a waterway which is usually constructed, and with a uniform section. We will use the more international English convention, that such a waterway is called a canal, and we will use the words ”waterway”, ”stream”, or ”channel” as generic terms which can describe any type of irregular river or regular canal or sewer with a free surface. 1.1 Types of channel flow to be studied (a) Steady uniform flow dn (b) Steady gradually-varied flow dn Normal depth (c) Steady rapidly-varied flow (d) Unsteady flow Figure 1-1. Different types of flow in an open channel Case (a) – Steady uniform flow: Steady flow is where there is no change with time, ∂/∂t ≡ 0. Distant from control structures, gravity and friction are in balance, and if the cross-section is constant, the flow is uniform, ∂/∂x ≡ 0. We will examine empirical laws which predict flow for given bed slope and roughness and channel geometry. Case (b) – Steady gradually-varied flow: Gravity and friction are in balance here too, but when a control is introduced which imposes a water level at a certain point, the height of the surface varies along the channel for some distance. For this case we will develop the differential equation which describes how conditions vary along the waterway. Case (c) – Steady rapidly-varied flow: Figure 1-1(c) shows three separate gradually-varied flow states separated by two rapidly-varied regions: (1) flow under a sluice gate and (2) a hydraulic jump. The complete problem as presented in the figure is too difficult for us to study, as the basic hydraulic approximation that variation is gradual and that the pressure distribution is hydrostatic breaks down in the rapid transitions between the different gradually-varied states. We can, however, analyse such problems by considering each of the almost-uniform flow states and consider energy or momentum conservation between them as appropriate. In these sorts of problems we will assume that the slope of the stream 4
  • 5. Open channel hydraulics John Fenton balances the friction losses and we treat such problems as frictionless flow over a generally-horizontal bed, so that for the individual states between rapidly-varied regions we usually consider the flow to be uniform and frictionless, so that the whole problem is modelled as a sequence of quasi-uniform flow states. Case (d) – Unsteady flow: Here conditions vary with time and position as a wave traverses the waterway. We will obtain some results for this problem too. 1.2 Properties of channel flow z = η y z min z = z Figure 1-2. Cross-section of flow, showing isovels, contours on which velocity normal to the section is constant. Consider a section of a waterway of arbitrary section, as shown in Figure 1-2. The x co-ordinate is horizontal along the direction of the waterway (normal to the page), y is transverse, and z is vertical. At the section shown the free surface is z = η, which we have shown to be horizontal across the section, which is a good approximation in many flows. 1.2.1 Discharge across a cross-section The volume flux or discharge Q at any point is Q = Z A u dA = UA where u is the velocity component in the x or downstream direction, and A is the cross-sectional area. This equation defines the mean horizontal velocity over the section U . In most hydraulic applications the discharge is a more important quantity than the velocity, as it is the volume of water and its rate of propagation, the discharge, which are important. 1.2.2 A generalisation – net discharge across a control surface Having obtained the expression for volume flux across a plane surface where the velocity vector is normal to the surface, we introduce a generalisation to a control volume of arbitrary shape bounded by a control surface CS. If u is the velocity vector at any point throughout the control volume and ˆn is a unit vector with direction normal to and directed outwards from a point on the control surface, then u · ˆn on the control surface is the component of velocity normal to the control surface. If dS is an elemental area of the control surface, then the rate at which fluid volume is leaving across the control surface over that 5
  • 6. Open channel hydraulics John Fenton elemental area is u · ˆndS, and integrating gives Total rate at which fluid volume is leaving across the control surface = Z CS u · ˆndS. (1.1) If we consider a finite length of channel as shown in Figure 1-3, with the control surface made up of u1 nˆ1 u2 nˆ 2 Figure 1-3. Section of waterway and control surface with vertical ends the bed of the channel, two vertical planes across the channel at stations 1 and 2, and an imaginary enclosing surface somewhere above the water level, then if the channel bed is impermeable, u · n ˆ= 0 there; u = 0 on the upper surface; on the left (upstream) vertical plane u · n ˆ= −u1, where u1 is the horizontal component of velocity (which varies across the section); and on the right (downstream) vertical plane u · n ˆ= +u2. Substituting into equation (1.1) we have Z Total rate at which fluid volume is leaving across the control surface = − A1 u1 dA + Z A2 u2 dA = −Q1 + Q2. If the flow is steady and there is no increase of volume inside the control surface, then the total rate of volume leaving is zero and we have Q1 = Q2. While that result is obvious, the results for more general situations are not so obvious, and we will generalise this approach to rather more complicated situations – notably where the water surface in the Control Surface is changing. 1.2.3 A further generalisation – transport of other quantities across the control surface We saw that u · ˆndS is the volume flux through an elemental area – if wemultiply by fluid density ρ then ρ u · ˆndS is the rate at which fluid mass is leaving across an elemental area of the control surface, with a corresponding integral over the whole surface. Mass flux is actually more fundamental than volume flux, for volume is not necessarily conserved in situations such as compressible flow where the density varies. However in most hydraulic engineering applications we can consider volume to be conserved. Similarly we can compute the rate at which almost any physical quantity, vector or scalar, is being transported across the control surface. For example, multiplying the mass rate of transfer by the fluid velocity u gives the rate at which fluid momentum is leaving across the control surface, ρuu · ˆndS. 1.2.4 The energy equation in integral form for steady flow Bernoulli’s theorem states that: In steady, frictionless, incompressible flow, the energy per unit mass p/ρ+gz +V 2/2 is constant 6
  • 7. Open channel hydraulics John Fenton along a streamline, where V is the fluid speed, V 2 = u2+v2+w2, inwhich (u, v, w) are velocity components in a cartesian co-ordinate system (x, y, z) with z vertically upwards, g is gravitational acceleration, p is pressure and ρ is fluid density. In hydraulic engineering it is usually more convenient to divide by g such that we say that the head p/ρg + z + V 2/2g is constant along a streamline. In open channel flows (and pipes too, actually, but this seems never to be done) we have to consider the situation where the energy per unit mass varies across the section (the velocity near pipe walls and channel boundaries is smaller than in the middle while pressures and elevations are the same). In this case we cannot apply Bernoulli’s theorem across streamlines. Instead, we use an integral form of the energy equation, although almost universally textbooks then neglect variation across the flow and refer to the governing theorem as ”Bernoulli”. Here we try not to do that. The energy equation in integral form can be written for a control volume CV bounded by a control surface CS, where there is no heat added or work done on the fluid in the control volume: ∂ ∂t Z CV ρ e dV | {z } Rate at which energy is increasing inside the CV + Z (p + ρe) u.ˆndS | {z } CS Rate at which energy is leaving the CS = 0, (1.2) where t is time, ρ is density, dV is an element of volume, e is the internal energy per unit mass of fluid, which in hydraulics is the sum of potential and kinetic energies e = gz + 1 2 ¡ u2 + v2 + w2¢ , where the velocity vector u = (u, v, w) in a cartesian coordinate system (x, y, z) with x horizontally along the channel and z upwards, ˆn is a unit vector as above, p is pressure, and dS is an elemental area of the control surface. Here we consider steady flow so that the first term in equation (1.2) is zero. The equation becomes: Z CS ³ p + ρgz + ρ 2 ¡ u2 + v2 + w2¢´ u.ˆn dS = 0. We intend to consider problems such as flows in open channels where there is usually no important contribution from lateral flows so that we only need to consider flow entering across one transverse face of the control surface across a pipe or channel and leaving by another. To do this we have the problem of integrating the contribution over a cross-section denoted by A which we also use as the symbol for the cross-sectional area. When we evaluate the integral over such a section we will take u to be the velocity along the channel, perpendicular to the section, and v and w to be perpendicular to that. The contribution over a section of area A is then ±E, where E is the integral over the cross-section: E = Z A ³ p + ρgz + ρ 2 ¡ u2 + v2 + w2¢´ u dA, (1.3) and we take the ± depending on whether the flow is leaving/entering the control surface, because u.ˆn = ±u. In the case of no losses, E is constant along the channel. The quantity ρQE is the total rate of energy transmission across the section. Now we consider the individual contributions: (a) Velocity head term ρ 2 R A ¡ u2 + v2 + w2 ¢ u dA If the flow is swirling, then the v and w components will contribute, and if the flow is turbulent there will be extra contributions as well. It seems that the sensible thing to do is to recognise that all velocity components and velocity fluctuations will be of a scale given by the mean flow velocity in the stream at 7
  • 8. Open channel hydraulics John Fenton that point,and so we simply write, for the moment ignoring the coefficient ρ/2: Z A ¡ u2 + v2 + w2¢ u dA = αU 3A = α Q3 A2 , (1.4) which defines α as a coefficient which will be somewhat greater than unity, given by α = R A ¡ u2 + v2 + w2 ¢ u dA U 3A . (1.5) Conventional presentations define it as being merely due to the non-uniformity of velocity distribution across the channel: α = R A u3 dA U 3A , however we suggest that is more properly written containing the other velocity components (and turbu-lent contributions as well, ideally). This coefficient is known as a Coriolis coefficient, in honour of the French engineer who introduced it. Most presentations of open channel theory adopt the approximation that there is no variation of velocity over the section, such that it is assumed that α = 1, however that is not accurate. Montes (1998, p27) quotes laboratory measurements over a smooth concrete bed giving values of α of 1.035-1.064, while for rougher boundaries such as earth channels larger values are found, such as 1.25 for irrigation canals in southern Chile and 1.35 in the Rhine River. For compound channels very much larger values may be encountered. It would seem desirable to include this parameter in our work, which we will do. (b) Pressure and potential head terms These are combined as Z A (p + ρgz) u dA. (1.6) The approximation we now make, common throughout almost all open-channel hydraulics, is the ”hy-drostatic approximation”, that pressure at a point of elevation z is given by p ≈ ρg × height of water above = ρg (η − z) , (1.7) where the free surface directly above has elevation η. This is the expression obtained in hydrostatics for a fluid which is not moving. It is an excellent approximation in open channel hydraulics except where the flow is strongly curved, such as where there are short waves on the flow, or near a structure which disturbs the flow. Substituting equation (1.7) into equation (1.6) gives ρg Z A η u dA, for the combination of the pressure and potential head terms. If we make the reasonable assumption that η is constant across the channel the contribution becomes ρgη Z A u dA = ρgηQ, from the definition of discharge Q. (c) Combined terms Substituting both that expression and equation (1.4) into (1.3) we obtain E = ρgQ μ η + α 2g Q2 A2 ¶ , (1.8) 8
  • 9. Open channel hydraulics John Fenton which, in the absence of losses, would be constant along a channel. This energy flux across entry and exit faces is that which should be calculated, such that it is weighted with respect to the mass flow rate. Most presentations pretend that one can just apply Bernoulli’s theorem, which is really only valid along a streamline. However our results in the end are not much different. We can introduce the concept of the Mean Total Head H such that H = Energy flux g × Mass flux = E g × ρQ = η + α 2g Q2 A2 , (1.9) which has units of length and is easily related to elevation in many hydraulic engineering applications, relative to an arbitrary datum. The integral version, equation (1.8), is more fundamental, although in common applications it is simpler to use the mean total head H, which will simply be referred to as the head of the flow. Although almost all presentations of open channel hydraulics assume α = 1, we will retain the general value, as a better model of the fundamentals of the problem, which is more accurate, but also is a reminder that although we are trying to model reality better, its value is uncertain to a degree, and so are any results we obtain. In this way, it is hoped, we will maintain a sceptical attitude to the application of theory and ensuing results. (d) Application to a single length of channel – including energy losses We will represent energy losses by ΔE. For a length of channel where there are no other entry or exit points for fluid, we have Eout = Ein − ΔE, giving, from equation (1.8): ρQout μ gη + α 2 Q2 A2 ¶ out = ρQin μ gη + α 2 Q2 A2 ¶ in − ΔE, and as there is no mass entering or leaving, Qout = Qin = Q, we can divide through by ρQ and by g, as is common in hydraulics: μ η + α 2g Q2 A2 ¶ out = μ η + α 2g Q2 A2 ¶ in − ΔH, where we have written ΔE = ρgQ × ΔH, where ΔH is the head loss. In spite of our attempts to use energy flux, as Q is constant and could be eliminated, in this head form the terms appear as they are used in conventional applications appealing to Bernoulli’s theorem, but with the addition of the α coefficients. 2. Conservation of energy in open channel flow In this section and the following one we examine the state of flow in a channel section by calculating the energy and momentum flux at that section, while ignoring the fact that the flow at that section might be slowly changing. We are essentially assuming that the flow is locally uniform – i.e. it is constant along the channel, ∂/∂x ≡ 0. This enables us to solve some problems, at least to a first, approximate, order. We can make useful deductions about the behaviour of flows in different sections, and the effects of gates, hydraulic jumps, etc.. Often this sort of analysis is applied to parts of a rather more complicated flow, such as that shown in Figure 1-1(c) above, where a gate converts a deep slow flow to a faster shallow flow but with the same energy flux, and then via an hydraulic jump the flow can increase dramatically in depth, losing energy through turbulence but with the same momentum flux. 2.1 The head/elevation diagram and alternative depths of flow Consider a steady (∂/∂t ≡ 0) flow where any disturbances are long, such that the pressure is hydro-static. We make a departure from other presentations. Conventionally (beginning with Bakhmeteff in 1912) they introduce a co-ordinate origin at the bed of the stream and introduce the concept of ”specific energy”, which is actually the head relative to that special co-ordinate origin. We believe that the use of 9
  • 10. Open channel hydraulics John Fenton that datum somehow suggests that the treatment and the results obtained are special in some way. Also, for irregular cross-sections such as in rivers, the ”bed” or lowest point of the section is poorly defined, and we want to minimise our reliance on such a point. Instead, we will use an arbitrary datum for the head, as it is in keeping with other areas of hydraulics and open channel theory. Over an arbitrary section such as in Figure 1-2, from equation (1.9), the head relative to the datum can be written H = η + αQ2 2g 1 A2(η) , (2.1) where we have emphasised that the cross-sectional area for a given section is a known function of surface elevation, such that we write A(η). A typical graph showing the dependence of H upon η is shown in Figure 2-1, which has been drawn for a particular cross-section and a constant value of discharge Q, such that the coefficient αQ2/2g in equation (2.1) is constant. η1 ηc η2 zmin Hc Surface elevation η H = η + αQ2 Head H = E/ρgQ 2g 1 A2(η) H = η 1 2 Figure 2-1. Variation of head with surface elevation for a particular cross-section and discharge The figure has a number of important features, due to the combination of the linear increasing function η and the function 1/A2(η) which decreases with η. • In the shallow flow limit as η → zmin (i.e. the depth of flow, and hence the cross-sectional area A(η), both go to zero while holding discharge constant) the value of H ∼ αQ2/2gA2(η) becomes very large, and goes to ∞ in the limit. • In the other limit of deep water, as η becomes large, H ∼ η, as the velocity contribution becomes negligible. • In between these two limits there is a minimum value of head, at which the flow is called critical flow, where the surface elevation is ηc and the head Hc. • For all other H greater than Hc there are two values of depth possible, i.e. there are two different flow states possible for the same head. • The state with the larger depth is called tranquil, slow, or sub-critical flow, where the potential to make waves is relatively small. • The other state, with smaller depth, of course has faster flow velocity, and is called shooting, fast, or super-critical flow. There is more wave-making potential here, but it is still theoretically possible for the flow to be uniform. • The two alternative depths for the same discharge and energy have been called alternate depths. 10
  • 11. Open channel hydraulics John Fenton That terminology seems to be not quite right – alternate means ”occur or cause to occur by turns, go repeatedly from one to another”. Alternative seems better - ”available as another choice”, and we will use that. • In the vicinity of the critical point, where it is easier for flow to pass from one state to another, the flow can very easily form waves (and our hydrostatic approximation would break down). • Flows can pass from one state to the other. Consider the flow past a sluice gate in a channel as shown in Figure 1-1(c). The relatively deep slow flow passes under the gate, suffering a large reduction in momentum due to the force exerted by the gate and emerging as a shallower faster flow, but with the same energy. These are, for example, the conditions at the points labelled 1 and 2 respectively in Figure 2-1. If we have a flow with head corresponding to that at the point 1 with surface elevation η1 then the alternative depth is η2 as shown. It seems that it is not possible to go in the other direction, from super-critical flow to sub-critical flow without some loss of energy, but nevertheless sometimes it is necessary to calculate the corresponding sub-critical depth. The mathematical process of solving either problem, equivalent to reading off the depths on the graph, is one of solving the equation αQ2 2gA2(η1) + η1 | {z } H1 = αQ2 2gA2(η2) + η2 | {z } H2 (2.2) for η2 if η1 is given, or vice versa. Even for a rectangular section this equation is a nonlinear tran-scendental equation which has to be solved numerically by procedures such as Newton’s method. 2.2 Critical flow δη δA B Figure 2-2. Cross-section of waterway with increment of water level We now need to find what the condition for critical flow is, where the head is a minimum. Equation (2.1) is H = η + α 2g Q2 A2(η) , and critical flow is when dH/dη = 0: dH dη = 1− αQ2 gA3(η) × dA dη = 0. The problem now is to evaluate the derivative dA/dη. From Figure 2-2, in the limit as δη → 0 the element of area δA = B δη,such that dA/dη = B, the width of the free surface. Substituting, we have the condition for critical flow: α Q2B gA3 = 1. (2.3) 11
  • 12. Open channel hydraulics John Fenton This can be rewritten as α (Q/A)2 g (A/B) = 1, and as Q/A = U , the mean velocity over the section, and A/B = D, the mean depth of flow, this means that Critical flow occurs when α U 2 gD = 1, that is, when α × (Mean velocity)2 g × Mean depth = 1. (2.4) We write this as αF 2 = 1 or √αF = 1, (2.5) where the symbol F is the Froude number, defined by: F = Q/A p gA/B = U √gD = Mean velocity √g × Mean depth. The usual statement in textbooks is that ”critical flow occurs when the Froude number is 1”. We have chosen to generalise this slightly by allowing for the coefficient α not necessarily being equal to 1, giving αF 2 = 1at critical flow. Any form of the condition, equation (2.3), (2.4) or (2.5) can be used. The mean depth at which flow is critical is the ”critical depth”: Dc = α U 2 g = α Q2 gA2 . (2.6) 2.3 The Froude number The dimensionless Froude number is traditionally used in hydraulic engineering to express the relative importance of inertia and gravity forces, and occurs throughout open channel hydraulics. It is relevant where the water has a free surface. It almost always appears in the form of αF 2 rather than F . It might be helpful here to define F by writing F 2 = Q2B gA3 . Consider a calculation where we attempt to quantify the relative importance of kinetic and potential energies of a flow – and as the depth is the only vertical scale we have we will use that to express the potential energy. We write Mean kinetic energy per unit mass Mean potential energy per unit mass = 12 αU 2 gD 2 αF 2, = 1 which indicates something of the nature of the dimensionless number αF 2. Flows which are fast and shallow have large Froude numbers, and those which are slow and deep have small Froude numbers. For example, consider a river or canal which is 2m deep flowing at 0.5ms−1 (make some effort to imagine it - we can well believe that it would be able to flow with little surface disturbance!). We have F = U √gD ≈ 0.5 √10 × 2 = 0.11 and F 2 = 0.012 , and we can imagine that the rough relative importance of the kinetic energy contribution to the potential contribution really might be of the order of this 1%. Now consider flow in a street gutter after rain. The velocity might also be 0.5ms−1, while the depth might be as little as 2 cm. The Froude number is F = U √gD ≈ 0.5 √10 × 0.02 = 1.1 and F 2 = 1.2 , 12
  • 13. Open channel hydraulics John Fenton which is just super-critical, and we can easily imagine it to have many waves and disturbances on it due to irregularities in the gutter. It is clear that αF 2 expresses the scale of the importance of kinetic energy to potential energy, even if not in a 1 : 1 manner (the factor of 1/2). It seems that αF 2 is a better expression of the relative importance than the traditional use of F . In fact, we suspect that as it always seems to appear in the form αF 2 = αU 2/gD, we could define an improved Froude number, Fimproved = αU 2/gD, which explicitly recognises (a) that U 2/gD is more fundamental than U/√gD, and (b) that it is the weighted value of u2 over the whole section, αU 2, which better expresses the importance of dynamic contributions. However, we will use the traditional definition F = U/√gD. In tutorials, assignments and exams, unless advised otherwise, you may assume α = 1, as has been almost universally done in textbooks and engineering practice. However we will retain α as a parameter in these lecture notes, and we recommend it also in professional practice. Retaining it will, in general, give more accurate results, but also, retaining it while usually not being quite sure of its actual value reminds us that we should not take numerical results as accurately or as seriously as we might. Note that, in the spirit of this, we might well use g ≈ 10 in practical calculations! Rectangular channel There are some special simple features of rectangular channels. These are also applicable to wide chan-nels, where the section properties do not vary much with depth, and they can be modelled by equivalent rectangular channels, or more usually, purely in terms of a unit width. We now find the conditions for critical flow in a rectangular section of breadth b and depth h. We have A = bh. From equation (2.3) the condition for critical flow for this section is: αQ2 gb2h3 = 1, (2.7) but as Q = U bh, this is the condition αU 2 gh = 1. (2.8) Some useful results follow if we consider the volume flow per unit width q: q = Q b = U bh b = U h. (2.9) Eliminating Q from (2.7) or U from (2.8) or simply using (2.6) with Dc = hc for the rectangular section gives the critical depth, when H is a minimum: hc = μ α q2 g ¶1/3 . (2.10) This shows that the critical depth hc for rectangular or wide channels depends only on the flow per unit width, and not on any other section properties. As for a rectangular channel it is obvious and convenient to place the origin on the bed, such that η = h. Then equation (2.1) for critical conditions when H is a minimum, H = Hc becomes Hc = hc + α 2g Q2 A2c = hc + α 2g Q2 b2h2c = hc + αq2 2g 1 h2c , and using equation (2.10) to eliminate the q2 term: Hc = hc + h3c 2 1 h2c = 3 2 hc or, hc = 2 3 Hc. (2.11) 2.4 Water level changes at local transitions in channels Now we consider some simple transitions in open channels from one bed condition to another. 13
  • 14. Open channel hydraulics John Fenton Sub-critical flow over a step in a channel or a narrowing of the channel section: Consider the 1 2 Δ Figure 2-3. Subcritical flow passing over a rise in the bed Surface elevation η ¾ Head H = E/ρgQ Critical constriction Constriction Upstream section 1 2 4 3 2’ Figure 2-4. Head/Surface-elevation relationships for three cross-sections flow as shown in Figure 2-3. At the upstream section the (H, η) diagram can be drawn as indicated in Figure 2-4. Now consider another section at an elevation and possible constriction of the channel. The corresponding curve on Figure 2-4 goes to infinity at the higher value of zmin and the curve can be shown to be pushed to the right by this raising of the bed and/or a narrowing of the section. At this stage it is not obvious that the water surface does drop down as shown in Figure 2-3, but it is immediately explained if we consider the point 1 on Figure 2-4 corresponding to the initial conditions. As we assume that no energy is lost in travelling over the channel constriction, the surface level must be as shown at point 2 on Figure 2-4, directly below 1 with the same value of H, and we see how, possibly against expectation, the surface really must drop down if subcritical flow passes through a constriction. Sub-critical flow over a step or a narrowing of the channel section causing critical flow: Consider 14
  • 15. Open channel hydraulics John Fenton now the case where the step Δ is high enough and/or the constriction narrow enough that the previously sub-critical flow is brought to critical, going from point 1 as before, but this time going to point 2’ on Figure 2-4. This shows that for the given discharge, the section cannot be constricted more than this amount which would just take it to critical. Otherwise, the (H, η) curve for this section would be moved further to the right and there would be no real depth solutions and no flow possible. In this case the flow in the constriction would remain critical but the upstream depth would have to increase so as to make the flow possible. The step is then acting as a weir, controlling the flow such that there is a unique relationship between flow and depth. Super-critical flow over a step in a channel or a narrowing of the channel section: Now consider super-critical flow over the same constriction as shown in Figure 2-5. In this case the depth actually increases as the water passes over the step, going from 3 to 4, as the construction in Figure 2-4 shows. Δ 3 4 Figure 2-5. Supercritical flow passing over a hump in the bed. Themathematical problem in each of these cases is to solve an equation similar to (2.2) for η2, expressing the fact that the head is the same at the two sections: αQ2 2gA21 (η1) + η1 | {z } H1 = αQ2 2gA22 (η2) + η2 | {z } H2 . (2.12) As the relationship between area and elevation at 2 is different from that at 1, we have shown two different functions for area as a function of elevation, A1(η1) and A2(η2). Example: A rectangular channel of width b1 carries a flow of Q, with a depth h1. The channel section is narrowed to a width b2 and the bed raised by Δ, such that the flow depth above the bed is now h2. Set up the equation which must be solved for h2. Equation (2.12) can be used. If we place the datum on the bed at 1, then η1 = h1 and A1(η1) = b1η1 = b1h1. Also, η2 = Δ + h2 and A2(η2) = b2 (η2 − Δ) = b2h2. The equation becomes αQ2 2gb21 h21 + h1 = αQ2 2gb22 h22 + Δ + h2, to be solved for h2, OR, αQ2 2gb21 h21 + h1 = αQ2 2gb22 (η2 − Δ)2 + η2, to be solved for η2. In either case the equation, after multiplying through by h2 or η2 respectively, becomes a cubic, which has no simple analytical solution and generally has to be solved numerically. Below we will present methods for this. 2.5 Some practical considerations 2.5.1 Trapezoidal sections 15
  • 16. Open channel hydraulics John Fenton γ 1 B h W Figure 2-6. Trapezoidal section showing important quantities Most canals are excavated to a trapezoidal section, and this is often used as a convenient approximation to river cross-sections too. In many of the problems in this course we will consider the case of trapezoidal sections. We will introduce the terms defined in Figure 2-6: the bottom width is W , the depth is h, the top width is B, and the batter slope, defined to be the ratio of H:V dimensions is γ. From these the following important section properties are easily obtained: Top width : B = W +2γh Area : A = h (W + γh) p 1 + γ2h, Wetted perimeter : P = W + 2 where we will see that the wetted perimeter is an important quantity when we consider friction in chan-nels. (Ex. Obtain these relations). 2.5.2 Solution methods for alternative depths Here we consider the problem of solving equation (2.12) numerically: αQ2 2gA21 (η1) + η1 = αQ2 2gA22 (η2) + η2, where we assume that we know the upstream conditions at point 1 and we have to find η2. The right side shows sufficiently complicated dependence on η2 that even for rectangular sections we have to solve this problem numerically. Reference can be made to any book on numerical methods for solving nonlinear equations, but here we briefly describe some techniques and then develop a simplified version of a robust method 1. Trial and error - evaluate the right side of the equation with various values of η2 until it agreeswith the left side. This is simple, but slow to converge and not suitable for machine computation. 2. Direct iteration - re-arrange the equation in the form η2 = H1 − αQ2 2gA22 (η2) and successively evaluate the right side and substitute for η2. We can show that this converges only if the flow at 2 is subcritical (αF 2 < 1), the more common case. Provided one is aware of that limitation, the method is simple to apply. 3. Bisection - choose an initial interval in which it is known a solution lies (the value of the function changes sign), then successively halve the interval and determine in which half the solution lies each time until the interval is small enough. Robust, not quite as simply programmed, but will always converge to a solution. 4. Newton’s method - make an estimate and then make successively better ones by travelling down the local tangent. This is fast, and reliable if a solution exists. We write the equation to be solved as f (η2) = η2 + αQ2 2gA22 (η2) − H1 (= 0 when the solution η2 is found). (2.13) 16
  • 17. Open channel hydraulics John Fenton Then, if η(n) 2 is the nth estimate of the solution, Newton’s method gives a better estimate: η(n+1) 2 = η(n) 2 − f (η(n) 2 ) f 0(η(n) 2 ) , (2.14) where f 0(h2) = ∂f /∂η2. In our case, from (2.13): f 0(η2) = ∂f (η2) ∂η2 = 1− αQ2 gA32 (η2) ∂A2 ∂η2 = 1− αQ2B2(η2) gA(η2) 32 = 1− αF 2(η2), which is a simple result - obtained using the procedure we used for finding critical flow in an arbitrary section. Hence, the procedure (2.14) is η(n+1) 2 = η(n) 2 − η(n) 2 + αQ2 2gA22 2 ) −H1 (η(n) 1−αF (n)2 2 . (2.15) Note that this will not converge as quickly if the flow at 2 is critical, where both numerator and denominator go to zero as the solution is approached, but the quotient is still finite. This expression looks complicated, but it is simple to implement on a computer, although is too complicated to appear on an examination paper in this course. These methods will be examined in tutorials. 2.6 Critical flow as a control - broad-crested weirs For a given discharge, the (H, η) diagram showed that the bed cannot be raised or the section narrowed more than the amount which would just take it to critical. Otherwise there would be no real depth solutions and no flow possible. If the channel were constricted even more, then the depth of flow over the raised bed would remain constant at the critical depth, and the upstream depth would have to increase so as to make the flow possible. The step is then acting as a weir, controlling the flow. hc hc hc Figure 2-7. A broad-crested weir spillway, showing the critical depth over it providing a control. Consider the situation shown in Figure 2-7 where the bed falls away after the horizontal section, such as on a spillway. The flow upstream is subcritical, but the flow downstream is fast (supercritical). Some-where between the two, the flow depth must become critical - the flow reaches its critical depth at some point on top of the weir, and the weir provides a control for the flow, such that a relationship between flow and depth exists. In this case, the head upstream (the height of the upstream water surface above the sill) uniquely determines the discharge, and it is enough to measure the upstream surface elevation where the flow is slow and the kinetic part of the head negligible to provide a point on a unique relationship between that head over the weir and the discharge. No other surface elevation need be measured. Figure 2-8 shows a horizontal flow control, a broad-crested weir, in a channel. In recent years there has been a widespread development (but not in Australia, unusually) of such broad-crested weirs placed in streams where the flow is subcritical both before and after the weir, but passes through critical on the 17
  • 18. Open channel hydraulics John Fenton weir. There is a small energy loss after the flume. The advantage is that it is only necessary to measure the upstream head over the weir. Small energy loss hc hc hc Figure 2-8. A broad-crested weir 3. Conservation of momentum in open channel flow 3.1 Integral momentum theorem P Control volume M1 1 2 M2 Figure 3-1. Obstacle in stream reducing the momentum flux We have applied energy conservation principles. Now we will apply momentum. We will consider, like several problems above, relatively short reaches and channels of prismatic (constant) cross-section such that the small contributions due to friction and the component of gravity down the channel are roughly in balance. Figure 3-1 shows the important horizontal contributions to force and momentum in the channel, where there is a structure applying a force P to the fluid in the control volume we have drawn. The momentum theorem applied to the control volume shown can be stated: the net momentum flux leaving the control volume is equal to the net force applied to the fluid in the control volume. The momentum flux is defined to be the surface integral over the control surface CS: Z CS (p ˆn + u ρu.ˆn) dS, where ˆn is a unit vector normal to the surface, such that the pressure contribution on an element of area dS is the force p dS times the unit normal vector ˆn giving its direction; u is the velocity vector such that u.ˆn is the component of velocity normal to the surface, u.ˆn dS is the volume rate of flow across the surface, multiplying by density gives the mass rate of flow across the surface ρu.ˆn dS, and multiplying by velocity gives uρu.ˆn dS, the momentum rate of flow across the surface. We introduce i, a unit vector in the x direction. On the face 1 of the control surface in Figure 3-1, as the outwards normal is in the upstream direction, we have ˆn = −i, and u = u1i, giving u.ˆn = −u1and the 18
  • 19. Open channel hydraulics John Fenton vector momentum flux across face 1 is M1 = −i R A1 ¡ p1 + ρu21 ¢ dA = −i M1, where the scalar quantity M1 = R A1 ¡ p1 + ρu21 ¢ dA. Similarly, on face 2 of the control surface, as the outwards normal is in the downstream direction, we have ˆn = i and u = u2i, giving u.ˆn = +u2 and the vector momentum flux across face 2 is M2 = +i R A2 ¡ p2 + ρu22 ¢ dA = +i M2 with scalar quantity M2 = R A2 ¡ p2 + ρu22 ¢ dA. Using the momentum theorem, and recognising that the horizontal component of the force of the body on the fluid is −P i, then we have, writing it as a vector equation but including only x (i) components: M1 +M2 = −P i AsM1 = −M1i andM2 = +M2i, we can write it as a scalar equation giving: P = M1 − M2, (3.1) where P is the force of the water on the body (or bodies). 3.1.1 Momentum flux across a section of channel From the above, it can be seen how useful is the concept of the horizontal momentum flux at a section of the flow in a waterway: M = Z A ¡ p + ρu2¢ dA. We attach different signs to the contributions depending on whether the fluid is leaving (+ve) or en-tering (-ve) the control volume. As elsewhere in these lectures on open channel hydraulics we use the hydrostatic approximation for the pressure: p = ρg(η − z), which gives M = ρ Z A ¡ g(η − z) + u2¢ dA. Now we evaluate this in terms of the quantities at the section. Pressure and elevation contribution ρ R A g(η − z) dA : The integral R A (η − z) dA is simply the first moment of area about a transverse horizontal axis at the surface, we can write it as R A (η − z) dA = A¯h, (3.2) where ¯h is the depth of the centroid of the section below the surface. Velocity contribution ρ R A u2 dA : Now we have the task of evaluating the square of the horizontal velocity over the section. As with the kinetic energy integral, it seems that the sensible thing to do is to recognise that all velocity components and velocity fluctuations will be of a scale given by the mean flow velocity in the stream at that point, and so we simply write Z A u2 dA = βU 2A = β Q2 A , (3.3) 19
  • 20. Open channel hydraulics John Fenton which defines β as a coefficient which will be somewhat greater than unity, given by β = R A u2 dA U 2A . (3.4) This coefficient is known as a Boussinesq coefficient, in honour of the French engineer who introduced it, who did much important work in the area of the non-uniformity of velocity and the non-hydrostatic nature of the pressure distribution. Most presentations of open channel theory adopt the approximation that there is no variation of velocity over the section, such that it is assumed that β = 1. Typical real values are β = 1.05 − 1.15, somewhat less than the Coriolis energy coefficient α. Combining: We can substitute to give the expression we will use for the Momentum Flux: M = ρ ¡ gA¯h + βU 2A ¢ = ρ μ gA¯h + β Q2 A ¶ = ρg μ A(h)¯ h(h) + βQ2 g 1 A(h) ¶ (3.5) where we have shown the dependence on depth in each term. This expression can be compared with that for the head as defined in equation (2.1) but here expressed relative to the bottom of the channel: H = h + αQ2 2g 1 A2(h) . The variation with h is different between this and equation (3.5). For large h, H ∼ h, while M ∼ A(h) h(h), which for a rectangular section goes like h2. For small h, H ∼ 1/A2(h), and M ∼ 1/A(h). Note that we can re-write equation (3.5) in terms of Froude number (actually appearing as F 2 – yet again) to indicate the relative importance of the two parts, which we could think of as ”static” and ”dynamic” contributions: M = ρgA¯h μ 1+βF 2 A/B ¯h ¶ . The ratio (A/B) /¯ h, mean depth to centroid depth, will have a value typically of about 2. Example: Calculate (a) Head (using the channel bottom as datum) and (b) Momentum flux, for a rectangular section of breadth b and depth h. We have A = bh, h = h/2. Substituting into equations (2.1) and (3.5) we obtain H = h + αQ2 2gb2 × 1 h2 and, M = ρ μ gb 2 × h2+ βQ2 b × 1 h ¶ . Note the quite different variation with h between the two quantities. 3.1.2 Minimum momentum flux and critical depth We calculate the condition for minimum M: ∂M ∂h = ∂ ∂h (A(h)¯ h(h)) − βQ2 g 1 A2(h) ∂A ∂h = 0. (3.6) The derivative of the first moment of area about the surface is obtained by considering the surface increased by an amount h + δh ∂(A¯ h) ∂h = lim δh→0 (A(h)¯ h(h))h+δh − A(h)¯ h(h)) δh . (3.7) The situation is as shown in Figure 3-2. The first moment of area about an axis transverse to the channel 20
  • 21. Open channel hydraulics John Fenton Depth to centroid of hatched area: δh/2 B h Depth to centroid of white area: h +δh δh Figure 3-2. Geometrical interpretation of calculation of position of centroid at the new surface is: (A(h)¯ h(h))h+δh = A(h) × (¯h + δh) + B × δh × δh 2 , so that, substituting into equation (3.7), in the limit δh → 0, ∂(A¯ h) ∂h = lim δh→0 A(h) × (¯h + δh) + B × δh × δh/2 − A(h)¯ h(h)) δh = A(h) = A, (3.8) which is surprisingly simple. Substituting both this and ∂A/∂h = B in equation (3.6), we get the condition for minimum M: βQ2B gA3 = βF 2 = 1, (3.9) which is a similar condition for the minimum energy, but as in general α6= β, the condition for minimum momentum is not the same as that for minimum energy. 3.1.3 Momentum flux -depth diagram If the cross-section changes or there are other obstacles to the flow, the sides of the channel and/or the obstacles will also exert a force along the channel on the fluid. We can solve for the total force exerted between two sections if we know the depth at each. In the same way as we could draw an (H, η) diagram for a given channel section, we can draw an (M, η) diagram. It is more convenient here to choose the datum on the bed of the channel so that we can interpret the surface elevation η as the depth h. Figure 3-3 shows a momentum flux – depth (M, h) diagram. Note that it shows some of the main features of the (H, h) diagram, with two possible depths for the same momentum flux – called conjugate depths. However the limiting behaviours for small and large depths are different for momentum, compared with energy. 3.2 Flow under a sluice gate and the hydraulic jump Consider the flow problem shown at the top of Figure 3-4, with sub-critical flow (section 0) controlled by a sluice gate. The flow emerges from under the gate flowing fast (super-critically, section 1). There has been little energy loss in the short interval 0-1, but the force of the gate on the flow has substantially reduced its momentum flux. It could remain in this state, however here we suppose that the downstream level is high enough such that a hydraulic jump occurs, where there is a violent turbulent motion and in a short distance the water changes to sub-critical flow again. In the jump there has been little momentum loss, but the turbulence has caused a significant loss of energy between 1-2. After the jump, at stage 2, the flow is sub-critical again. We refer to this depth as being sequent to the original depth. In the bottom part of Figure 3-4 we combine the (H, h) and (M, h) diagrams, so that the vertical axis is 21
  • 22. Open channel hydraulics John Fenton Momentum flux M Depth h h1 h2 3 2 1 4 Rectangular: M ~ h2 Rectangular: M ~ 1/ h P h4 P h3 Figure 3-3. Momentum flux – depth diagram, showing effects of a momentum loss P for subcritical and supercrit-ical flow. 0 h 0 P Head H, Momentum flux M Depth h Head H Momentum flux M 0 2 2 1 1 H 0 = H1 M 1 = M 2 M 0 2 h 1 h P 0 1 2 Figure 3-4. Combined Head and Momentum diagrams for the sluice gate and hydraulic jump problem 22
  • 23. Open channel hydraulics John Fenton depth h and the two horizontal axes are head H and momentum flux M, with different scales. We now outline the procedure we follow to analyse the problem of flow under a sluice gate, with upstream force P , and a subsequent hydraulic jump. • We are given the discharge Q and the upstream depth h0, and we know the cross-sectional details of the channel. • We can compute the energy and momentum at 0, H0 and M0 (see points 0 on the M −H −h plot). • As energy is conserved between 0 and 1, the depth h1 can be calculated by solving the energy equation with H1 = H0, possibly using Newton’s method. • In fact, this depth may not always be realisable, if the gate is not set at about the right position. The flow at the lip of the gate leaves it vertically, and turns around to horizontal, so that the gate opening must be larger than h1. A rough guide is that the gate opening must be such that h1 ≈ 0.6×Gate opening. • With this h1 we can calculate the momentum flux M1. • The force on the gate P (assuming that the channel is prismatic) can be calculated from: P = M0 − M1 = ρ μ gAh+β Q2 A ¶ 0 − ρ μ gAh + β Q2 A ¶ 1 • Across the hydraulic jump momentum is conserved, such that M2 = M1: μ gAh+β Q2 A ¶ 2 = μ gAh+β Q2 A ¶ 1 • This gives a nonlinear equation for h2 to be solved numerically (note that A and¯h are both functions of h). In the case of a rectangular channel the equation can be written 1 2 gh21 + βq2 h1 = 1 2 gh22 + βq2 h2 , where q = Q/b, the discharge per unit width. In fact it can be solved analytically. Grouping like terms on each side and factorising: (h2 − h1)(h2 + h1) = 2βq2 g μ 1 h1 − 1 h2 ¶ , h22 h1 + h21h2 − 2βq2 g = 0, which is a quadratic in h2, with solutions h2 = − h1 2 ± s h21 4 + 2βq2 gh1 , but we cannot have a negative depth, and so only the positive sign is taken. Dividing through by h1: h2 h1 = − 1 2 + s 1 4 + 2βq2 gh31 = 1 2 μq 1 + 8βF 2 ¶ 1 −1 • Sometimes the actual depth of the downstream flow is determined by the boundary condition further downstream. If it is not deep enough the actual jump may be an undular hydraulic jump, which does not dissipate as much energy, with periodic waves downstream. • The pair of depths (h1, h2) for which the flow has the same momentum are traditionally called the conjugate depths. 23
  • 24. Open channel hydraulics John Fenton • The loss in energy H2 − H1 can be calculated. For a rectangular channel it can be shown that ΔH = H1 − H2 = (h2 − h1)3 4h1h2 . 3.3 The effects of streams on obstacles and obstacles on streams 3.3.1 Interpretation of the effects of obstacles in a flow Slow (sub-critical) approach flow Figure 3-5 shows that the effect of a drag force is to lower the P P h 1 h 2 h 1 2 M c h Figure 3-5. Effect of obstacles on a subcritical flow water surface (counter-intuitive!?) if the flow is slow (sub-critical). Fast (super-critical) approach flow Figure 3-6 shows that the effect of a drag force on a super-critical P h 2 P h 1 h 1 2 M c h Figure 3-6. Effect of obstacles on a supercritical flow flow is to raise the water surface. In fact, the effect of the local force only spreads gradually through the stream by turbulent diffusion, and the predicted change in cross-section will apply some distance downstream where the flow has become uniform (rather further than in the diagrams here). A practical example is the fast flow downstream of a spillway, shown in Figure 3-7, where the flow becomes subcritical via a hydraulic jump. If spillway blocks are used, the water level downstream need not be as high, possibly with large savings in channel construction. 3.3.2 Bridge piers - slow approach flow Consider flow past bridge piers as shown in Figure 3-8. As the bridge piers extend throughout the flow, for the velocity on the pier we will take the mean upstream velocity V = Q/A1, and equation (3.14) can 24
  • 25. Open channel hydraulics John Fenton P 2 P h 2* 1 2 * 2 h = depth without blocks 1 h 1 2 M Figure 3-7. Effect of spillway blocks on lowering the water level in a spillway pool Plan Side elevation c h 1 2 P c h 1 2 M Figure 3-8. Flow past bridge piers and their effect on the flow be used. 3.3.3 Flow in a narrowing channel - choked flow We consider cases where the width reduction is more than in a typical bridge pier problem, such that the flow in the throat may become critical, the throat becomes a control, and the flow is said to be choked. If so, the upstream depth is increased, to produce a larger momentum flux there so that the imposed force due to the convergence now just produces critical flow in the throat. In problems such as these, it is very helpful to remember that for a rectangular section, equation (2.10): hc = ¡ αq2/g ¢1/3 , or, re-written, q = p gh3c /α, where q = Q/b, the flow per unit width, and also to observe that at critical depth, equation (2.11): H = hc + αQ2 2gb2h2c = hc + αq2 2gh2c = 3 2 hc, so that hc = 2 3 H. It is clear that by reducing b, q = Q/b is increased, until in this case, criticality is reached. While 25
  • 26. Open channel hydraulics John Fenton c h Elevation c h Plan Figure 3-9. Flow through contraction sufficiently narrow that it becomes critical generally this is not a good thing, as the bridge would then become a control, where there is a relationship between flow and depth, this becomes an advantage in flow measurement applications. In critical flow flumes only an upstream head is needed to calculate the flow, and the structure is deliberately designed to bring about critical depth at the throat. One way of ensuring this is by putting in a rise in the bed at the throat. Note that in the diagram the critical depth on the hump is greater than that upstream because the width has been narrowed. 3.3.4 Drag force on an obstacle As well as sluices and weirs, many different types of obstacles can be placed in a stream, such as the piers of a bridge, blocks on the bed, Iowa vanes, the bars of a trash-rack etc. or possibly more importantly, the effects of trees placed in rivers (”Large Woody Debris”), used in their environmental rehabilitation. It might be important to know what the forces on the obstacles are, or in flood studies, what effects the obstacles have on the river. Substituting equation (3.5) into equation (3.1) (P = M1 − M2) gives the expression: P = ρ μ gA¯h + β Q2 A ¶ 1 − ρ μ gA¯h + β Q2 A ¶ 2 , (3.10) so that if we know the depth upstream and downstream of an obstacle, the force on it can be calculated. Usually, however, the calculation does not proceed in that direction, as one wants to calculate the effect of the obstacle on water levels. The effects of drag can be estimated by knowing the area of the object measured transverse to the flow, a, the drag coefficient Cd, and V , themean fluid speed past the object: P = 1 2 ρCdV 2a, (3.11) and so, substituting into equation (3.10) gives, after dividing by density, 1 2 CdV 2a = μ gA¯h + β Q2 A ¶ 1 − μ gA¯h + β Q2 A ¶ 2 . (3.12) We will write the velocity V on the obstacle as being proportional to the upstream velocity, such that we write V 2 = γd μ Q A1 ¶2 , (3.13) 26
  • 27. Open channel hydraulics John Fenton where γd is a coefficient which recognises that the velocity which impinges on the object is generally not equal to the mean velocity in the flow. For a small object near the bed, γd could be quite small; for an object near the surface it will be slightly greater than 1; for objects of a vertical scale that of the whole depth, γd ≈ 1. Equation (3.12) becomes 1 2 γd Cd Q2 A21 a = μ gA¯h + β Q2 A ¶ 1 − μ gA¯h + β Q2 A ¶ 2 (3.14) A typical problem is where the downstream water level is given (sub-critical flow, so that the control is downstream), and we want to know by how much the water level will be raised upstream if an obstacle is installed. As both A1 and h1 are functions of h1, the solution is given by solving this transcendental equation for h1. In the spirit of approximation which can be used in open channel hydraulics, and in the interest of simplicity and insight, we now obtain an approximate solution. 3.3.5 An approximate method for estimating the effect of channel obstructions on flooding Momentum flux M Depth h 1 h 3 2 1 4 4 h P 3 h Tangent to (M,h) curve Approximate h1 Exact h1 2 h Figure 3-10. Momentum flux – depth diagram showing the approximate value of d1 calculated by approximating the curve by its tangent at 2. Now an approximation to equation (3.14) will be obtained which enables a direct calculation of the change in water level due to an obstacle, without solving the transcendental equation. We consider a linearised version of the equation, which means that locally we assume a straight-line approximation to the momentum diagram, for a small reduction in momentum, as shown in Figure 3-10. Consider a small change of surface elevation δh going from section 1 to section 2, and write the expres-sion for the downstream area A2 = A1 + B1δh. It has been shown above (equation 3.8) that ∂(A¯ h) ∂h = A, and so we can write an expression for A2¯h 2 in terms of A1¯h 1 and the small change in surface elevation: A2¯h 2 = A1¯h 1 + δh ∂(A¯ h) ∂h ¯¯¯¯ 1 = A1¯h 1 + δh A1, 27
  • 28. Open channel hydraulics John Fenton and so equation (3.14) gives us, after dividing through by g: 1 2 γd Cd Q2 gA21 a = −δh A1 + β Q2 gA1 − β Q2 g (A1 + B1δh) = −δh A1 + β Q2 gA1 Ã 1 − μ 1 + B1 A1 ¶ −1 δh ! . Now we use a power series expansion in δh to simplify the last term, neglecting terms like (δh)2. For ε small, (1 + ε)−1 ≈ 1 − ε, and so 1 2 γd Cd Q2 gA21 a ≈ −δh A1 + β Q2B1 gA21 δh. We can now solve this to give an explicit approximation for δh: δh ≈ 12 γd Cd Q2 gA31 a β Q2B1 gA31 − 1 . It is simpler to divide both sides by the mean depth A1/B1 to give: δh A1/B1 = 1 2 γd Cd F 2 1 a A1 βF 2 1 − 1 . We do not have to worry here that for subcritical flow we do not necessarily know the conditions at point 1, but instead we know them at the downstream point 2. Within our linearising approximation, we can use either the values at 1 or 2 in this expression, and so we generalise by dropping the subscripts altogether, so that we write δh A/B = 1 2 γd Cd F 2 a A βF 2 − 1 = 1 2 γd Cd a A × F 2 βF 2 − 1 . (3.15) Thus we see that the relative change of depth (change of depth divided by mean depth) is directly proportional to the coefficient of drag and the fractional area of the blockage, as we might expect. The result is modified by a term which is a function of the square of the Froude number. For subcritical flow the denominator is negative, and so is δh, so that the surface drops, as we expect, and as can be seen when we solve the problem exactly using the momentum diagram. If upstream is supercritical, the surface rises. Clearly, if the flow is near critical (βF 2 1 ≈ 1) the change in depth will be large (the gradient on the momentum diagram is vertical), when the theory will have limited validity. Example: In a proposal for the rehabilitation of a river it is proposed to install a number of logs (”Large Woody Debris” or ”Engineered Log Jam”). If a single log of diameter 500mm and 10m long were placed transverse to the flow, calculate the effect on river height. The stream is roughly 100m wide, say 10m deep in a severe flood, with a drag coefficient Cd ≈ 1. The all-important velocities are a bit uncertain. We might assume a mean velocity of say 6ms−1, and velocity on the log of 2ms−1. Assume β = 1.1. We have the values A = 100 × 10 = 1000m2, a = 0.5 × 10 = 5m2 F 2 = U 2/gD = 62/10/10 = 0.36, γd = 22/62 ≈ 0.1 and substituting into equation (3.15) gives δh A/B ≈ 1 2 γdCd a A1 β − 1 F 2 1 = 1 2 × 0.1 × 1 × 5 1000 1.1 − 1 0.36 = −1.5 × 10−4, so that multiplying by the mean depth, δη = −1.5 × 10−4 × 10 = −1.5mm. The negative value is the 28
  • 29. Open channel hydraulics John Fenton change as we go downstream, thus we see that the flow upstream is raised by 1.5mm. 4. Uniform flow in prismatic channels Uniform flow is where the depth does not change along the waterway. For this to occur the channel properties also must not change along the stream, such that the channel is prismatic, and this occurs only in constructed canals. However in rivers if we need to calculate a flow or depth, it is common to use a cross-section which is representative of the reach being considered, and to assume it constant for the application of this theory. 4.1 Features of uniform flow and relationships for uniform flow • There are two forces in balance in steady flow: – The component of gravity downstream along the channel, and – the shear stress at the sides which offers resistance to the flow, which increases with flow veloc-ity. • If a channel is long and prismatic (slope and section do not change) then far from the effects of controls the two can be in balance, and if the flow is steady, the mean flow velocity and flow depth remain constant along the channel, giving uniform flow, at normal depth. A L P τ0 τ0 θ g sin θ g Figure 4-1. Slice of uniform channel flow showing shear forces and body forces per unit mass acting Consider a slice of uniform flow in a channel of length L and cross-sectional area A, as shown in Figure 4-1. The component of gravity force along the channel is ρ × AL × g sin θ, where θ is the angle of inclination of the channel, assumed positive downwards. The shear force is τ 0 × L × P , where τ 0 is the shear stress, and P is the wetted perimeter of the cross-section. As the two are in balance for uniform flow, we obtain τ 0 ρ = g A P sin θ. Now, τ 0/ρ has units of velocity squared; we combine g and the coefficient relating the mean 29
  • 30. Open channel hydraulics John Fenton velocity U at a section to that velocity, giving Chézy’s law (1768): U = C p RS0, where C is the Chézy coefficient (with units L1/2T−1), R = A/P is the hydraulic radius (L), and S0 = sinθ is the slope of the bed, positive downwards. The tradition in engineering is that we use the tangent of the slope angle, so this is valid for small slopes such that sin θ ≈ tan θ. • However there is experimental evidence that C depends on the hydraulic radius in the form C ∼ R1/6 (Gauckler, Manning), and the law widely used is Manning’s Law: U = 1 n R2/3S1/2 0 , where n is the Manning coefficient (units of L−1/3T), which increases with increasing roughness. Typical values are: concrete - 0.013, irrigation channels - 0.025, clean natural streams - 0.03, streams with large boulders - 0.05, streams with many trees - 0.07. Usually the units are not shown. • Multiplying by the area, Manning’s formula gives the discharge: Q = U A = 1 n A5/3 P 2/3 p S0, (4.1) in which both A and P are functions of the flow depth. Similarly, Chézy’s law gives Q = C A3/2 P 1/2 p S0. (4.2) Both equations show how flow increases with cross-sectional area and slope and decreases with wetted perimeter. 4.2 Computation of normal depth If the discharge, slope, and the appropriate roughness coefficient are known, either of equations (4.1) and (4.2) is a transcendental equation for the normal depth hn, which can be solved by the methods described earlier. We can gain some insight and develop a simple scheme by considering a trapezoidal cross-section, where the bottom width is W , the depth is h, and the batter slopes are (H:V) γ : 1 (see Figure 2-6). The following properties are easily shown to hold (the results have already been presented above): Top width B W +2γh Area A h(W + γh) Wetted perimeter P W +2 p 1 + γ2h In the case of wide channels, (i.e. channels rather wider than they are deep, h ¿ W , which is a common case) the wetted perimeter does not show a lot of variation with depth h. Similarly in the expression for the area, the second factor W +γh (the mean width) does not show a lot of variation with h either – most of the variation is in the first part h. Hence, if we assume that these properties hold for cross-sections of a more general nature, we can rewrite Manning’s law: Q = 1 n A5/3(h) P 2/3(h) p S0 = √S0 n (A(h)/h)5/3 P 2/3(h) × h5/3, where most of the variation with h is contained in the last term h5/3, and by solving for that term we can re-write the equation in a form suitable for direct iteration h = μ Qn √S0 ¶3/5 × P 2/5(h) A(h)/h , 30
  • 31. Open channel hydraulics John Fenton where the first term on the right is a constant for any particular problem, and the second term is expected to be a relatively slowly-varying function of depth, so that the whole right side varies slowly with depth – a primary requirement that the direct iteration scheme be convergent and indeed be quickly convergent. Experience with typical trapezoidal sections shows that this works well and is quickly convergent. How-ever, it also works well for flow in circular sections such as sewers, where over a wide range of depths the mean width does not vary much with depth either. For small flows and depths in sewers this is not so, and a more complicated method might have to be used. Example: Calculate the normal depth in a trapezoidal channel of slope 0.001, Manning’s coef-ficient n = 0.04, width 10m, with batter slopes 2 : 1, carrying a flow of 20m3 s−1. We have A = h (10 + 2 h), P = 10+4.472 h, giving the scheme h = μ Qn √S0 ¶3/5 × (10 + 4.472 h)2/5 10 + 2 h = 6.948 × (10 + 4.472 h)2/5 10 + 2 h and starting with h = 2 we have the sequence of approximations: 2.000, 1.609, 1.639, 1.637 - quite satisfactory in its simplicity and speed. 4.3 Conveyance It is often convenient to use the conveyance K which contains all the roughness and cross-section prop-erties, such that for steady uniform flow Q = K p S0, such that, using an electrical analogy, the flow (current) is given by a ”conductance” (here conveyance) multiplied by a driving potential, which, here in this nonlinear case, is the square root of the bed slope. In more general non-uniform flows below we will see that we use the square root of the head gradient. With this definition, if we use Manning’s law for the flow, K is defined by K = 1 n × A μ A P ¶2/3 = 1 n × A5/3 P 2/3 , (4.3) where K is a function of the roughness and the local depth and cross-section properties. Textbooks often use conveyance to provide methods for computing the equivalent conveyance of compound sections such as that shown in Figure 4-2. However, for such cases where a river has overflowed its banks, the flow situation is much more likely to be more two-dimensional than one-dimensional. The extent of the various elemental areas and the Manning’s roughnesses of the different parts are all such as to often render a detailed ”rational” calculation unjustified. 2 3 1 Figure 4-2. In the compound channel in the figure, even though the surface might actually be curved as shown and 31
  • 32. Open channel hydraulics John Fenton the downstream slope and/or bed slope might be different across the channel, the tradition is that we assume it to be the same. The velocities in the individual sections are, in general, different. We write Manning’s law for each section based on the mean bed slope: Q1 = K1S1/2 0 , Q2 = K2S1/2 0 , Q3 = K3S1/2 0 In a general case with n sub-sections, the total discharge is Q = Xn i=1 Qi = Xn i=1 KiS1/2 0 = S1/2 0 Xn i=1 Ki = S1/2 0 K where we use the symbol K for the total conveyance: K = Xn i=1 Ki = Xn i=1 A5/3 i niP 2/3 i . 5. Steady gradually-varied non-uniform flow Steady gradually-varied flow is where the conditions (possibly the cross-section, but often just the sur-face elevation) vary slowly along the channel but do not change with time. The most common situation where this arises is in the vicinity of a control in a channel, where there may be a structure such as a weir, which has a particular discharge relationship between the water surface level and the discharge. Far away from the control, the flow may be uniform, and there the relationship between surface elevation and discharge is in general a different one, typically being given by Manning’s law, (4.1). The transition between conditions at the control and where there is uniform flow is described by the gradually-varied flow equation, which is an ordinary differential equation for the water surface height. The solution will approach uniform flow if the channel is prismatic, but in general we can treat non-prismatic waterways also. In sub-critical flow the flow is relatively slow, and the effects of any control can propagate back up the channel, and so it is that the numerical solution of the gradually-varied flow equation also proceeds in that direction. On the other hand, in super-critical flow, all disturbances are swept downstream, so that the effects of a control cannot be felt upstream, and numerical solution also proceeds downstream from the control. Solution of the gradually-varied flow equation is a commonly-encountered problem in open channel hydraulics, as it is used to determine, for example, how far upstream water levels might be increased, and hence flooding enhanced, due to downstream works, such as the installation of a bridge. 5.1 Derivation of the gradually-varied flow equation Consider the elemental section of waterway of length Δx shown in Figure 5-1. We have shown stations 1 and 2 in what might be considered the reverse order, but we will see that for the more common sub-critical flow, numerical solution of the governing equation will proceed back up the stream. Considering stations 1 and 2: Total head at 2 = H2 Total head at 1 = H1 = H2 − HL, and we introduce the concept of the friction slope Sf which is the gradient of the total energy line such that HL = Sf × Δx. This gives H1 = H2 − Sf Δx, 32
  • 33. Open channel hydraulics John Fenton U 2 / 2 g Total energy line α 2 U 2 / 2 g α 1 2 h S f Δx S0 Δx 2 1 1 h Sub-critical flow Δx Figure 5-1. Elemental section of waterway and if we introduce the Taylor series expansion for H1: H1 = H2 + Δx dH dx + . . . , substituting and taking the limit Δx → 0 gives dH dx = −Sf , (5.1) an ordinary differential equation for the head as a function of x. To obtain the frictional slope, we use either of the frictional laws of Chézy orManning (or a smooth-wall formula), where we make the assumption that the equation may be extended from uniform flow (where the friction slope equals the constant bed slope) to this non-uniform case, such that the discharge at any point is given by, for the case of Manning: Q = 1 n A5/3 P 2/3 p Sf , but where we have used the friction slope Sf rather than bed slope S0, as in uniform flow. Solving for Sf : the friction slope is given by Sf = Q2 K2(h) , (5.2) where we have used the conveyance K, which was defined in equation (4.3), but we repeat here, K (h) = 1 n A5/3 P 2/3 , showing the section properties to be a function of the local depth, where we have restricted our attention to prismatic channels on constant slope. This now means that for a given constant discharge we can write the differential equation (5.1) as dH dx = −Sf (h). (5.3) As we have had to use local depth on the right side, we have to show the head to be a function of depth h, so that we write H = h + zmin + α 2g Q2 A2(h) . (5.4) 33
  • 34. Open channel hydraulics John Fenton Differentiating: dH dx = dh dx + dzmin dx − α g Q2 A3(h) dA(h) dx . (5.5) The derivative dzmin/dx = −S0, where S0 is the bed slope,which we have defined to be positive for the usual case of a downwards-sloping channel. Now we have to express the dA(h)/dx in terms of other quantities. In our earlier work we saw that if the surface changed by an amount Δh, then the change in area due to this was ΔA = B Δh, and so we can write dA(h)/dx = B dh/dx, and substituting these results into equation (5.5) gives dH dx = −S0 + μ 1 − α g Q2B(h) A3(h) ¶ dh dx = −S0 + ¡ 1 − αF 2(h) ¢ dh dx , where the Froude number has entered, shown here as a function of depth. Finally, substituting into (5.3) we obtain dh dx = S0 − Sf (h) 1 − αF 2(h) = S0 − Q2/K2(h) 1 − αF 2(h) , (5.6) a differential equation for depth h as a function of x, where on the right we have shown the functional dependence of the various terms. This, or the less-explicit form (5.3), are forms of the gradually-varied flow equation, from which a number of properties can be inferred. 5.2 Properties of gradually-varied flow and the governing equation • The equation and its solutions are important, in that they tell us how far the effects of a structure or works in or on a stream extend upstream or downstream. • It is an ordinary differential equation of first order, hence one boundary condition must be supplied to obtain the solution. In sub-critical flow, this is the depth at a downstream control; in super-critical flow it is the depth at an upstream control. • In general that boundary depth is not equal to the normal depth, and the differential equation de-scribes the transition from the boundary depth to normal depth – upstream for sub-critical flow, downstream for supercritical flow. The solutions look like exponential decay curves, and below we will show that they are, to a first approximation. • If that approximation is made, the resulting analytical solution is useful in providing us with some insight into the quantities which govern the extent of the upstream or downstream influence. • The differential equation is nonlinear, and the dependence on h is complicated, such that analytical solution is not possible without an approximation, and we will usually use numerical methods. • The uniform flow limit satisfies the differential equation, for when Sf = S0, dh/dx = 0, and the depth does not change. • As the flow approaches critical flow, when αF 2 → 1, then dh/dx → ∞, and the surface becomes vertical. This violates the assumption we made that the flow is gradually varied and the pressure distribution is hydrostatic. This is the one great failure of our open channel hydraulics at this level, that it cannot describe the transition between sub- and super-critical flow. 5.3 Classification system for gradually-varied flows The differential equation can be used as the basis for a dual classification system of gradually-varied flows: • one based on 5 conditions for slope, essentially as to how the normal depth compares with critical depth, and 3 conditions for the actual depth, and how it compares with both normal, and critical depths, as shown in the Table: 34
  • 35. Open channel hydraulics John Fenton Slope classification Steep slope: hn < hc Critical slope: hn = hc Mild slope: hn > hc Horizontal slope: hn = ∞ Adverse slope: hn does not exist Depth classification Zone 1: h > hn and hc Zone 2: h between hn and hc Zone 3: h < hn and hc Figure 5-2 shows the behaviour of the various solutions. In practice, the most commonly encountered are the M1, the backwater curve on a mild slope; M2, the drop-down curve on a mild slope, and S2, the drop-down curve on a steep slope. 5.4 Some practical considerations 5.4.1 Flood inundation studies Figure 5-3 shows a typical subdivision of a river and its flood plain for a flood inundation study, where solution of the gradually-varied flow equation would be required. It might be wondered how the present methods can be used for problems which are unsteady, such as the passage of a substantial flood, where on the front face of the flood wave the water surface is steeper and on the back face it is less steep. In many situations, however, the variation of the water slope about the steady slope is relatively small, and the wavelength of the flood is long, so that the steady model can be used as a convenient approxima-tion. The inaccuracies of knowledge of the geometry and roughness are probably such as to mask the numerical inaccuracies of the solution. Below we will present some possible methods and compare their accuracy. 5.4.2 Incorporation of losses It is possible to incorporate the losses due, say, to a sudden expansion or contraction of the channel, such as shown in Figure 5-4. After an expansion the excess velocity head is destroyed through turbulence. Before an expansion the losses will not be so large, but there will be some extra losses due to the convergence and enhanced friction. We assume that the expansion/contraction head loss can be written ΔHe = C μ Q2 2gA22 − Q2 2gA21 ¶ , where C ≈ 0.3 for expansions and 0.1 for a contraction. 5.5 Numerical solution of the gradually-varied flow equation Consider the gradually-varied flow equation (5.6) dh dx = S0 − Sf (h) 1 − αF 2(h) , where both Sf (h) = Q2/K2(h) and F 2(h) = Q2B(h)/gA3(h) are functions of Q as well as the depth h. However as Q is constant for a particular problem we do not show the functional dependence on it. The equation is a differential equation of first order, and to obtain solutions it is necessary to have a boundary condition h = h0 at a certain x = x0, which will be provided by a control. The problem may be solved using any of a number of methods available for solving ordinary differential equations which 35
  • 36. Open channel hydraulics John Fenton Figure 5-2. Typical gradually-varied flow surface profiles, drawn by Dr I. C. O’Neill. 36
  • 37. Open channel hydraulics John Fenton Typical cross-section used for 1-D analysis Edge of flood plain / Extent of 1% flood River banks Figure 5-3. Practical river problem with subdivision 2 1 Figure 5-4. Flow separation and head loss due to a contraction are described in books on numerical methods. These methods are usually accurate and can be found in many standard software packages. It is surprising that books on open channels do not recognise that the problem of numerical solution of the gradually-varied flow equation is actually a standard numerical problem, although practical details may make it more complicated. Instead, such texts use methods such as the ”Direct step method” and the ”Standard step method”. There are several software packages such as HEC-RAS which use such methods, but solution of the gradually-varied flow equation is not a difficult problem to solve for specific problems in practice if one knows that it is merely the solution of a differential equation, and here we briefly set out the nature of such schemes. The direction of solution is very important. If the different conventional cases in Figure 5-2 are exam-ined, it can be seen that for the mild slope (sub-critical flow) cases that the surface decays somewhat exponentially to normal depth upstream from a downstream control, whereas for steep slope (super-critical flow) cases the surface decays exponentially to normal depth downstream from an upstream control. This means that to obtain numerical solutions we will always solve (a) for sub-critical flow: from the control upstream, and (b) for super-critical flow: from the control downstream. 5.5.1 Euler’s method The simplest (Euler) scheme to advance the solution from (xi, hi) to (xi + Δxi, hi+1) is xi+1 ≈ xi + Δxi, where Δxi is negative for subcritical flow, (5.7) dh hi+1 ≈ hi + Δxi dx ¯¯¯¯ i = hi + Δxi S0 − Sf (hi) 1 − αF 2(hi) . (5.8) This is the simplest but least accurate of all methods – yet it might be appropriate for open channel 37
  • 38. Open channel hydraulics John Fenton problems where quantities may only be known approximately. One can use simple modifications such as Heun’s method to gain better accuracy – or even more simply, just take smaller steps Δxi. 5.5.2 Heun’s method In this case the value of hi+1 calculated from equation (5.8) is used as a first estimate h∗i+1, then the value of the right hand side of the differential equation is also calculated there, and the mean of the two values taken. That is, xi+1 ≈ xi + Δxi, again where Δxi is negative for subcritical flow, (5.9) h∗i+1 = hi + Δxi S0 − Sf (hi) 1 − αF 2(hi) , (5.10) hi+1 = hi + Δxi 2 μ S0 − Sf (hi) 1 − αF 2(hi) + S0 − Sf (h∗i+1) 1 − αF 2(h∗i+1) ¶ . (5.11) Neither of these two methods are presented in hydraulics textbooks as alternatives. Although they are simple and flexible, they are not as accurate as other less-convenient methods described further below. The step Δxi can be varied at will, to suit possible irregularly spaced cross-sectional data. 5.5.3 Predictor-corrector method – Trapezoidal method This is simply an iteration of the last method, whereby the step in equation (5.11) is repeated several times, at each stage setting h∗i+1 equal to the updated value of hi+1. This gives an accurate and conve-nient method, and it is surprising that it has not been used. 5.5.4 Direct step method Textbooks do present the Direct Step method, which is applied by taking steps in the height and calcu-lating the corresponding step in x. It is only applicable to problems where the channel is prismatic. The reciprocal of equation (5.1) is dx dH = − 1 Sf , which is then approximated by a version of Heun’s method, but which is not a correct rational approxi-mation: Δx = − ΔH ¯ Sf , (5.12) where a mean value of the friction slope is used. The procedure is: for the control point x0 and h0, calculate H0 from equation (5.4), then assume a finite value of depth change Δh to compute h1 = h0 + Δh, from which H1 is calculated from equation (5.4), giving ΔH = H1 − H0. Then with ¯ Sf = (Sf (h0)+Sf (h1))/2, equation (5.12) is used to calculate the corresponding Δx, giving x1 = x0 +Δx. The process is then repeated to give x2 and h2 and so on. It is important to choose the correct sign of Δh such that computations proceed in the right direction such that, for example, Δx is negative for sub-critical flow, and computations proceed upstream. The method has the theoretical disadvantage that it is an inconsistent approximation, in that it should actually be computing the mean of 1/Sf, namely 1/Sf , rather than 1/ ¯ Sf . More importantly it has practical disadvantages, such that it is applicable only to prismatic sections, results are not obtained at specified points in x, and as uniform flow is approached the Δx become infinitely large. However it is a surprisingly accurate method. 5.5.5 Standard step method This is an implicit method, requiring numerical solution of a transcendental equation at each step. It can be used for irregular channels, and is rather more general. In this case, the distance interval Δx is specified and the corresponding depth change calculated. In the Standard step method the procedure is 38
  • 39. Open channel hydraulics John Fenton to write ΔH = −Sf Δx, and then write it as H2(h2) − H1(h1) = − Δx 2 (Sf1 + Sf2) , for sections 1 and 2, where the mean value of the friction slope is used. This gives α Q2 2gA22 + Z2 + h2 = α Q2 2gA21 + Z1 + h1 − Δx 2 (Sf1 + Sf2) , where Z1 and Z2 are the elevations of the bed. This is a transcendental equation for h2, as this determines A2, P2, and Sf2. Solution could be by any of the methods we have had for solving transcendental equations, such as direct iteration, bisection, or Newton’s method. Although the Standard step method is an accurate and stable approximation, the lecturer considers it unnecessarily complicated, as it requires solution of a transcendental equation at each step. It would be much simpler to use a simple explicit Euler or Heun’s method as described above. Example: Consider a simple backwater problem to test the accuracy of the various methods. A trapezoidal channel with bottom width W = 10m, side batter slopes of 2:1, is laid on a slope of S0 = 10−4, and carries a flow of Q = 15m3 s−1. Manning’s coefficient is n = 0.025. At the downstream control the depth is 2.5m. Calculate the surface profile (and how far the effect of the control extends upstream). Use 10 computational steps over a length of 30km. 2.5 2.4 2.3 2.2 2.1 2 -45 -40 -35 -30 -25 -20 -15 -10 -5 0 Depth (m) x (km) Accurate solution Analytical approximation Trapezoidal Standard step Direct step Figure 5-5. Comparison of different solution methods – depth plotted. Figure 5-5 shows the results of the computations, where depth is plotted, while Figure 5-6 shows the same results, but where surface elevation is plotted, to show what the surface profile actually looks like. For relatively few computational points Euler’s method was not accurate, and neither was Heun’s method, and have not been plotted. The basis of accuracy is shown by the solid line, from a highly-accurate Runge-Kutta 4th order method. This is not recommended as a method, however, as it makes use of information from three intermediate points at each step, information which in non-prismatic channels is not available. It can be seen that the relatively simple Trapezoidal method is sufficiently accurate, certainly of acceptable practical accuracy. The Direct Step method was slightly more accurate, but the results show one of its disadvantages, that the distance between computational points becomes large as uniform flow is approached, and the points are at awkward distances. The last plotted point is at about −25km; using points closer to normal depth gave inaccurate results. The Standard Step method was very accurate, but is not plotted as it is complicated to apply. Of course, if more computational points 39
  • 40. Open channel hydraulics John Fenton 6.5 6 5.5 5 4.5 4 3.5 3 2.5 2 Accurate solution and normal depth Analytical approximation -45 -40 -35 -30 -25 -20 -15 -10 -5 0 Surface elevation (m) x (km) Trapezoidal Standard step Direct step Figure 5-6. Comparison of different solution methods – elevation plotted. were taken, more accurate results could be obtained. In this example we deliberately chose relatively few steps (10) so that the numerical accuracies of the methods could be compared. Also plotted on the figures is a dotted line corresponding to the analytical solution which will be de-veloped below. Although this was not as accurate as the numerical solutions, it does give a simple approximate result for the rate of decay and how far upstream the effects of the control extend. For many practical problems, this accuracy and simplicity may be enough. The channel dimensions are typical of a large irrigation canal in the Murray Valley - it is interesting that the effects of the control extend for some 30km! To conclude with a recommendation: the trapezoidal method, Heun’s method iterated several times is simple, accurate, and convenient. If, however, a simple approximate solution is enough, then the following analytical solution can be used. 5.6 Analytical solution Whereas the numerical solutions give us numbers to analyse, sometimes very few actual numbers are required, such as merely requiring how far upstream water levels are raised to a certain level, the effect of downstream works on flooding, for example. Here we introduce a different way of looking at a physical problem in hydraulics, where we obtain an approximate mathematical solution so that we can provide equations which reveal to us more of the nature of the problem than do numbers. Sometimes an understanding of what is important is more useful than numbers. Consider the water surface depth to be written h(x) = h0 + h1(x), where we use the symbol h0 for the constant normal depth, and h1(x) is a relatively small departure of the surface from the uniform normal depth. We use the governing differential equation (5.6) but we assume that the Froude number squared is sufficiently small that it can be ignored. This is not essential, but it makes the equations simpler to write and read. (As an example, consider a typical stream flowing at 0.5 m/s with a depth of 2m, giving F 2 = 0.0125 - there are many cases where F 2 can be neglected). 40
  • 41. Open channel hydraulics John Fenton The simplified differential equation can be written dh dx = S0 − S(h), where for purposes of simplicity we have dropped the subscript f on the friction slope, now represented by S. Substituting our expansion, we obtain dh1 dx = S0 − S(h0 + h1(x)). (5.13) Now we introduce the approximation that the h1 term is relatively small such that we can write for the friction term its Taylor expansion about normal flow: S(h0 + h1(x)) = S(h0) + h1(x) × dS dh (h0) + Terms proportional to h21 . We ignore the quadratic terms, write dS/dh(h0) as S/ 0 , and substituting into equation (5.13), we obtain dh1 dx = −S/ 0 h1 where we have used S(h0) = S0. This is an ordinary differential equation which we can solve analyt-ically. We have achieved this by ”linearising” about the uniform flow. Now, by separation of variables we can obtain the solution h1 = Ge−S/ 0 x, and the full solution is h = h0 + Ge−S/ 0 x, (5.14) where G is a constant which would be evaluated by satisfying the boundary condition at the control. This shows that the water surface is actually approximated by an exponential curve passing from the value of depth at the control to normal depth. In fact, we will see that as S/ 0 is negative, far upstream as x → −∞, the water surface approaches normal depth. Now we obtain an expression for S/ 0 in terms of the channel dimensions. From Manning’s law, S = n2Q2 P 4/3 A10/3 , and differentiating gives S/ = n2Q2 à 4 3 P 1/3 A10/3 dP dh − 10 3 P 4/3 A13/3 dA dh ! , which we can factorise, substitute dA/dh = B, and recognising the term outside the brackets, we obtain an analytical expression for the coefficient of x in the exponential function: 0 = n2Q2 P 4/3 −S/ 0 A10/3 0 μ 10 3 B A − 4 3 dP/dh P ¶¯¯¯¯ 0 = S0 μ 10 3 B0 A0 − 4 3 dP0/dh0 P0 ¶ . The larger this number, the more rapid is the decay with x. The formula shows that more rapid decay occurs with steeper slopes (large S0), smaller depths (B0/A0 = 1/D0, where D0 is the mean depth - if it decreases the overall coefficient increases), and smaller widths (P0 is closely related to width, the term involving it can be written d(log P0)/dh0: if P0 decreases the term decreases - relatively slowly - but the negative sign means that the effect is to increase the magnitude of the overall coefficient). Hence, generally the water surface approaches normal depth more quickly for steeper, shallower and narrower (i.e. steeper and smaller) streams. The free surface will decay to 10% of its original departure from 41
  翻译: