Ch2 Practical and Theoretical Issues in Channel Hydraulics

download Ch2 Practical and Theoretical Issues in Channel Hydraulics

of 78

description

Practical Channel Hydraulics

Transcript of Ch2 Practical and Theoretical Issues in Channel Hydraulics

  • Chapter 2

    Practical and theoretical issuesin channel hydraulics

    ABSTRACT

    This chapter is an introduction to the many practical and theoretical issues that needto be considered when attempting to model the flow of water in open channels andrivers. These issues are illustrated through several worked examples, as well as anumber of case studies on actual rivers. Consideration is also given to several relatedtopics, such as: how to schematise a river cross-section or reach, the difficulty ofestimating hydraulic roughness and resistance, the hydraulic characteristics of inbankand overbank flows, and flows through bridges and culverts. The data requirementsfor effective modelling are also considered.

    2.1 GETTING STARTED WITH SOME PRACTICALEXAMPLES ON CALCULATING FLOWSIN WATERCOURSES

    The flow of water in a river, water course or drainage system is governed by physicallaws related to the conservation of mass, energy and momentum. Their application toopen channel flow is the subject of this book. Although these physical laws and prin-ciples are sometimes difcult to apply to all types of fluid flow in river engineering, aswell as other areas in engineering and science, mainly due to the difculty of account-ing for energy degradation, some physical characteristics of open channel flow maybe readily understood and are surprising. In order to appreciate some key features ofopen channel flow, the following 5 examples are given to get started on some practicalproblems. They also help to introduce and define certain terms and parameters, as wellas illustrating certain issues addressed in later sections. Examples 1 to 3 are concernedwith a freely owing small-sized river or drainage channel, and Examples 4 & 5are concerned with flow in a reach where the channel has some hydraulic controls.Although these examples are straightforward for those familiar with hydraulic princi-ples and text books (e.g. Chadwick et al., 2004; Chanson, 1999; Chow, 1959; Cungeet al. 1980; Henderson, 1966), they should be worked through none-the-less as theyserve to illustrate some basic topics and issues in modelling mentioned later. Theyintroduce parameters such as specic energy, momentum, roughness, resistance toflow and energy losses, as well as the notation used in this book.

  • 16 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Example 1

    A river has a simple trapezoidal cross section with a bed width, b, of 15.0 m, sideslopes, s, of 1:1 (vertical : horizontal), a Mannings n of 0.030 and a longitudinalbed slope, So, of 3.0 104. Determine the discharge in the channel when thedepth of flow, h, is 1.5 m, using Mannings equation (1891), given as Eq. (2.1).Figures 2.1(a) & (b) illustrate typical river channels that might be considered to beapproximately trapezoidal in shape, but with different roughnesses.

    Q = AR2/3S1/2o /n (2.1)

    (a) low Manning n value (~0.03) (b) high Manning n value (~0.30)

    Figure 2.1 Two trapezoidal channels with different Manning n roughness values.

    Answer & comment

    Area, A = (b + sh)h = (15 + 1.0 1.5) 1.5 = 24.75 m2.Wetted perimeter, P = b + 2h(1 + s2) = 15 + 2 1.5(1 + 12) = 15 + 32 =19.243 m.Hydraulic radius, R = A/P = 24.75/19.243 = 1.286 m.Discharge, using Eq. (2.1), Q = 24.75 1.2862/3 0.00031/2/0.030 =16.90 m3s1.

    Note that the discharge is given explicitly by the Manning equation, because thedepth, h, is specified, and the whole cross section is used. For a long reach, witha constant slope and roughness, this depth is referred to as the normal depth andoften given the symbol hn. Since the whole section is used, this is essentially a singlechannel method of analysis, employing a simple one-dimensional (1-D) approachthat gives the discharge, but no other information about the velocity distributionwithin the channel (i.e. across the width or over the depth). For further details on thebackground to the Manning equation, see Yen (1991).

  • Practical and theoretical issues in channel hydraulics 17

    Example 2

    If the discharge in Example 1 is now reduced to 10 m3s1, determine the new depth offlow (normal depth), hn, and the critical depth, hc. Also determine the ratio of normalto critical depth, hn/hc, the Froude Number and whether the flow is subcritical orsupercritical.

    Answer & comment

    Equation (2.1) may be re-arranged with all known parameters on one side and all thegeometric parameters involving depth on the other side. Noting that R = A/P, then:

    Qn/S1/2o = A5/3/P2/3 (2.2)

    Since h occurs both in the area and wetted perimeter, whose constituent equationsdepend on the shape of the cross section, it is generally now not possible to obtain ananswer for the depth in explicit form using Mannings equation. Pivoting around thehighest power of h, a numerical method, such as fixed-point iteration, may be shownto give a new value for the normal depth, hj+1, in terms of a previous value, hj, as:

    hj+1 =[Qn

    S1/2o

    ]3/5 [b + 2hj1 + s2]2/5

    (b + shj) (2.3)

    Equation (2.3) may be solved numerically for hj+1, by assuming initially thathj = 0, and then by repeated iteration until hj+1 is sufficiently close to hj, givingultimately hn = 1.095 m. The normal depth of flow for the discharge of 10 m3s1 istherefore 1.095 m.

    Equation (2.3) may also be formulated into a dimensionless form, convenient forother purposes, as

    hj+1b

    =[

    Qn

    b8/3S1/2o

    ]3/5 [1 + (2hj/b)1 + s2]2/5

    (1 + shj/b) (2.4)

    Figure 2.2 shows the solution for h for flows in trapezoidal channels and circularculverts or pipes running part full, to illustrate the influence of each dimensionlessgroup. These types of flow are commonly found in engineering design and feature inthis book in relation to simple shaped rivers, culverts and drainage conduits.

    The critical depth, hc, is given when Fr = 1.0, where the Froude Number, Fr, isdefined by

    Fr = UgA/T

    (2.5)

    in which U = section mean velocity (= Q/A) and T = top width of channel (i.e. atthe free surface). At the critical depth and critical flow, Fr = 1, and the energy is

  • 18 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.2 Uniform depth relationships based on Eq. (2.4), and equivalent equation for circular culvertsrunning part full, showing influence of individual parameter groups.

    also a minimum, as may be deduced later from Eq. (2.12). The condition necessaryfor critical flow then follows from Eq. (2.5) and may be written in the general formfor channels of any cross sectional shape, as

    Q2TgA3

    = 1 (2.6)

    In order to calculate hc, a similar numerical procedure to the normal depthcalculation method is required. In this particular example, with a trapezoidal chan-nel, T = b + 2sh, which is a function of depth and therefore initially unknown.Re-arranging Eq. (2.6) for fixed-point iteration yields

    hj+1 =[Q2

    g

    ]1/3 [b + 2shj]1/3(b + shj) (2.7)

    This may also be put into a dimensionless form again, convenient for otherpurposes, as:

    hj+1b

    =[

    Qg1/2b5

    ]2/3 [1 + 2shj/b]1/3(1 + shj/b) (2.8)

    Using Eq. (2.7) iteratively, assuming initially again that hj = 0, gives hc = 0.354m.Thus hn/hc = 1.0952/0.3537 = 3.096 (i.e. hn > hc, therefore subcritical flow.See Fig. 2.12)

  • Practical and theoretical issues in channel hydraulics 19

    At the new normal depth, A = (b + sh)h = (15 + 1.0 1.0952) 1.0952 =17.628 m2.Top width, T = b + 2sh = 15 + 2 1.0 1.0952 = 17.190 m.U = Q/A = 10.0/17.6275 = 0.567 ms1.Fr = 0.5673/(9.807 17.6275/17.1904)1/2 = 0.179.Hence, since Fr < 1.0 the flow is confirmed as being subcritical.

    Note that an iterative approach is required to solve Eq. (2.1) because this is theinverse problem to Example 1, in which the discharge is given and not the depth. Sincehn > hc, this means that the flow is subcritical, a fact confirmed by the FroudeNumberbeing less than 1.0. The general form of the Froude Number, Eq. (2.5), should alwaysbe used in open channel work, and distinguished from the particular form adopted inchannels with a rectangular cross section. In that particular case, since A = bh andT = b, the Froude Number is given by

    Fr = Ugh

    (rectangular channel only) (2.9)

    and the critical depth may be readily obtained from Eq. (2.6) in the form

    hc = 3q2/g (rectangular channel only) (2.10)

    where the discharge per unit width, q, is defined by

    q = Q/b (2.11)The discharge per unit width is a useful concept, used widely in designing channel

    transitions, and also in a depth-averaged form of the Reynolds Averaged Navier-Stokes equations (RANS) which form the basis of the CES system described morefully in Chapter 3.

    Example 3

    If the river in Example 1 is re-graded, maintaining the same bed width and side slopes,but with a Mannings n = 0.025, what bed slope is required to convey 30 m3s1at a depth of 1.5 m? What will be the increase in water level for Q = 30 m3s1, ifsubsequently the roughness of the river reverts back to its original Mannings n value,with So remaining as calculated?

    Answer & comment

    A = 24.75 m2; P = 19.243 m; R = 1.286 m as before, then solving for slope:Bed slope, using Eq. (2.1), So = [30 0.025/(24.75 1.28622/3)] 2 = 6.565 104.For original n = 0.030, depth will increase. Using Eq. (2.3) gives hn = 1.673 m.Increase in water level = 0.173 m.

  • 20 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Note that with reduced roughness, but the same depth, the bed slope is increasedbecause the specified discharge has increased from 16.9 m3s1 to 30 m3s1. With theoriginal roughness, depth, new discharge but reduced bed slope, the depth is increased.This illustrates how useful the Manning equation is in relating the 4 basic parameters{Q, h, So, n} that commonly occur in many design problems in river engineering.Furthermore, Eq. (2.1) also shows how easy it is to calculate changes in any oneparameter that are governed by the other three.

    Example 4

    This example is based on the physical situation shown in Figure 2.3, and the cor-responding diagrams in Figure 2.4. A horizontal channel with a rectangular crosssection is 4.5 m wide and conveys 13.5 m3s1. A vertical sluice gate is incorporatedin the channel and the gate is regulated until the depth of flow just upstream is 3.0 m.Assuming no energy loss during the passage of water under the gate, find the depthof flow just downstream of the gate. Figure 2.3 shows a picture of this type of flow inthe laboratory, and Figure 2.4 shows the corresponding depth versus specic energy(h v E) and depth versus specic momentum (h v M) diagrams. The depth at section 2is known as the vena contracta (Latin for contracted vein), i.e. where the depth in thechannel is the lowest. The flow with a hydraulic jump is covered in Example 5.

    Answer & comment

    This example requires one to understand the concept of specific energy, E, defined by

    E = h + U2

    2g(2.12)

    Figure 2.3 Flow under a laboratory vertical sluice gate with a hydraulic jump downstream.

  • Practical and theoretical issues in channel hydraulics 21

    Figure 2.4 Schematic diagrams of specific energy and momentum for flow under a vertical sluice gatewith a hydraulic jump downstream.

    where E = specific energy per unit weight, = kinetic energy correction coefficient,taken here as 1.0 in order to simplify the problem. The technical definition of , andthe corresponding momentum correction coefficient, , are dealt with more fully inSection 2.2.5.

    Equation (2.12) may be developed for flow in a rectangular channel of breadth,b, with = 1.0, as a cubic equation in h

    E = h + q2

    2gh2(2.13)

    Thus with q = Q/b = 13.5/4.5 = 3.0 m3s1m1, then E1 = 3.0 + 3.02/(2 9.807 3.02) = 3.051 m.

    Assuming no energy loss across the sluice gate, as shown in Figure 2.4, thenE1 = E2 and the depth downstream of the gate, h2, may be obtained from

    E2 = h2 + q2

    2gh22

    Solving this for E2 = 3.051 m gives h2 = 0.417 m. Hence A2 = 4.5 0.4174 =1.878 m2 andU2 = 13.5/1.8783 = 7.187 ms1. Also, from Eq. (2.10), hc = 0.972 m,indicating that h2 < hc.

    Checking that the value of h2 gives the same specific energy, the kinetic head atsection 2 = U22/2g = 7.18742/(2 9.807) = 2.6337 m and hence E2 = 0.4174 +2.6337 = 3.0511 m (E1).

    Note that for a given specific energy there are generally two possible depths offlow, h1 and h2, known as alternate depths, one subcritical (Fr < 1) and the other

  • 22 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    supercritical (Fr > 1). These are denoted in Figure 2.4 by the two points 1 & 2 in theh v E diagram, where a vertical line for the given value of E intersects the curve fora given q. Note the value of U22/2g, and how the sluice gate changes the flow fromsubcritical to supercritical flow. The point on the h v E curve where the energy is aminimum, and where there is only one depth for a given energy, is where critical flowoccurs (i.e. Fr = 1.0). The depth there is known as the critical depth, hc, as alreadydefined by Eq. (2.6), and is associated usually with the transition from subcritical tosupercritical flow. It should also be noted that the downstream depth, h2, is seen tobe much smaller than the upstream depth h1 and is defined by h1 and q, not the gateopening, hg. The flow downstream is also confirmed to be supercritical since h2 < hc.The contraction coefficient, Cc (= h2/hg), may be calculated for the sluice gate if hg isknown. Typically, Cc is around 0.61 for such sluice gates. Alternatively, if hg, Cc andq are known, then the upstream depth, h1, may be calculated.

    Example 5

    If a hydraulic jump occurs downstream of the sluice gate in Example 4, as shown inFigures 2.3 & 2.4, determine the maximum tail water level which could be toleratedbefore the sluice gate becomes drowned out, i.e. when the downstream depth beginsto affect the depth of flow upstream of the gate. Also determine the height of the jumpand the energy and the power lost in the jump.

    Answer & comment

    The standard equation for a hydraulic jump in a rectangular channel (Chow, 1959) isas follows:

    h2h1

    = 12

    [(1 + 8Fr21) 1

    ](2.14)

    where h1 and h2 are the depths before and after the jump, known as conjugate depths.These should not be confused with the two alternate depths defined in Example 4 inrelation to specific energy. Applying Eq. (2.14) to sections 2 & 3, where section 2 isupstream of the jump and section 3 downstream, then:

    From Eq. (2.9), Fr2 = 7.1874/(9.807 0.4174)1/2 = 3.552.From Eq. (2.14), with appropriate notation changes, h3/h2 = 1/2 [(1 + 8 3.55242)1/2 1] = 4.549.Hence h3 = 4.5487 0.4174 = 1.899 m.Thus the height of the jump = h3 h2 = 1.8980.417 = 1.481 m.Themaximum tail water level to prevent drowning of the sluice gate is thus around1.90 m.

    In order to determine the energy loss, one needs to calculate E3.Downstream of the jump, U3 = Q/A3 = 13.5/(1.8986 4.5) = 1.580 ms1.

  • Practical and theoretical issues in channel hydraulics 23

    Hence E3 = 1.8986 + 1.58012/(2 9.807) = 2.026 m.The specific energy lost in the jump = E2 E3 = 3.051 2.026 = 1.025 m.This may also be determined from the standard relationship for a jump in a rect-

    angular channel (see Chow, 1959), where again h1 and h2 are the depths before andafter the jump

    E = (h2 h1)3/(4h1h2

    )(2.15)

    Thus between sections 2 & 3, E = (1.8986 0.4174)3/(4 0.4174 1.8986) =1.024 m.

    The power lost is given by

    P = gQE (2.16)

    Hence power loss= 10009.80713.51.0251 = 135713Watts {or 135.7 kW}.

    Note how the specific energy and momentum vary between sections 1, 2 & 3.The value of h3 in Figure 2.4 is seen to be much less than h1, indicating an energyloss somewhere between 1 & 3. Most of this loss is known to occur as a result ofthe intense mixing and turbulence within the hydraulic jump. Between 1 & 2, acrossthe sluice gate, it was therefore assumed that E = 0 and M = 0, whereas acrossthe hydraulic jump, between 2 & 3, it was assumed that E = 0 and M = 0.Since both energy and momentum feature later with regard to hydraulic structures,the momentum function, M, is now introduced. Applying Newtons 2nd law for flowin an open channel

    F1 F2 + F = Q (U2 U1) (2.17)

    where F1 & F2 are the hydrostatic forces acting on upstream and downstream crosssections (= gAz), z is the distance below the water surface to the centre of area ofthe cross section, F is any additional force acting on the fluid in the direction of flowand = momentum correction coefficient. Then in general for 1-D flow

    gA1z1 gA2z2 + F = Q (U2 U1) (2.18)

    Re-arranging terms in Eq. (2.18), dividing by g and assuming here for simplicitythat = 1, then

    {Q2

    gA1+ A1z1

    }

    {

    Q2

    gA2+ A2z2

    }

    + Fg

    = 0 (2.19)

    Defining the momentum function as

    M = Q2

    gA+ Az (2.20)

  • 24 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    then the force acting on the fluid in the flow direction between any two sections is

    Fg

    = M2 M1 (2.21)

    For the special case of a rectangular channel, with z = h/2, Eq. (2.20) shows that themomentum function is related to the specic force (= F/b, i.e. force per unit width) by

    M = b[q2

    gh+ h

    2

    2

    ]

    (2.22)

    Equation (2.22) has certain similarities with the Eq. (2.13) as already shown inFigure 2.4. Thus in Example 4, the force acting on the sluice gate, Fg, (= F) may becalculated from Eq. (2.21), giving

    M1 = 4.5 [32/(9.807 3.0) + 32/2] = 4.5 [0.3059 + 4.5] = 21.627 m3.M2 = 4.5 [32/(9.807 0.4174) + 0.41742/2] = 4.5 [2.1986 + 0.0871] =10.286 m3.F = 1000 9.807 (10.2858 21.6266) = 111218.98 N or 111.22 kN.Fg = F = +111.22 kN (i.e. force is positive and therefore acts in the direction offlow).

    It should be also noted that in the derivation of Eq. (2.14) it is assumed that thefrictional force, Ff , on bed under the short length of the jump is negligibly small. Thisis equivalent to stating that in Eq. (2.17) the force F = Ff = 0. However, in straightchannels where there may be blocks, or forward and rearward facing steps present (asshown in Figure 2.22), the relevant forces have to be specified in Eq. (2.17) in orderto produce an equation equivalent to Eq. (2.14). Where supercritical flow occurs inchannels with bends, care must be taken in dealing with the high super-elevation thatmight occur and any cross waves that may develop as a result of a hydraulic jumpforming in the channel. See Chow (1959) for further details on these topics.

    2.2 COMMON DIFFICULTIES IN MODELLING FLOWIN RIVERS AND WATERCOURSES

    The five examples in Section 2.1 were selected to illustrate simple flow behaviour inchannels where the geometry of the cross section is prismatic, i.e. a constant crosssectional shape in the flow direction, the roughness is uniformly distributed aroundthe wetted perimeter and characterised by the Manning equation, and there are nocomplex hydraulic structures, vegetation or sediment within the channel. In reality, inmost rivers and watercourses, all of these issues are generally of concern and must bemodelled appropriately.

    The previous pages and some of the examples have highlighted the three physicallaws related to the conservation of mass, energy and momentum. In each examplethe flow was considered to be steady, with no additional discharge being introducedwith distance along the channel, thus always ensuring mass conservation (Q1 = Q2).

  • Practical and theoretical issues in channel hydraulics 25

    The conservation of energy was assumed in the analysis of the vertical sluice gate(E1 = E2), allowing the depths either side to be determined, whereas the force actingon the fluid (or the reaction on the gate) was shown to be related to the change inmomentum (i.e. M1 = M2). In the case of the hydraulic jump, an energy loss occurred(i.e. E2 = E3), but momentum was conserved (M2 = M3). Dealing with energy lossesis, as shown later, one of the key issues to be considered in modelling flows in rivers.

    Although the same three physical principles outlined in Section 2.1 apply to flow innatural watercourses, the flow structure is typically three-dimensional and unsteady.The geometry of the river itself is also generally more complex and highly variable inthe streamwise direction. All of this makes the modelling of flow in rivers and water-courses more difcult. Furthermore, there is often little or no data concerning channelroughness or resistance, which are typically flow and season dependent. Many of thehydraulic structures that have been built in rivers are often outanked by the flow inflood conditions, or built to non-standard conditions, thus creating uncertainties inthe application or interpretation of various national or international codes of practice.There are also scale issues to consider when modelling at basin, reach, or cross-sectionscales, and even the schematisation of a short stretch of river requires careful consider-ation. Hydraulic controls may also vary with flow conditions, due to changing channelresistance, the behaviour of particular hydraulic structures or even human input. Forthese reasons, some of the common difculties and issues in modelling actual riverflows are now considered.

    2.2.1 Modelling f low in rivers and watercourses

    The use of models in investigating and solving practical river engineering problems isnow so commonplace, that it is easy sometimes to overlook their origin, nature andpurpose. There are different types of model that are now routinely used to analyseboth general flow behaviour patterns and also particular hydraulic issues, such as theextent of ooding, sediment deposition or erosion patterns, pollution incidents, vege-tation control and habitat biodiversity. A wide range of hydrodynamic tools are alsofrequently used in developing optimum design strategies for dealing with specic prac-tical problems. For example, these tools might include models for flood risk mapping,or models that simulate dispersion phenomena, sediment movement, bank erosion orgeneral flow behaviour. Such tools are typically based on a mathematical representa-tion of the physical processes (sometimes incorrectly referred to as a model), containadvanced numerical procedures, and employ complex computer techniques to carryout not only all the routine calculations, but also to present the results in a convenientform to the end-user, typically in graphical as well as numerical format.

    The origin of modern, computer-based, river modelling began in the 1960s andthe background to this is described byAbbot (1991), Cunge et al. (1980) andNakato&Ettema (1996). Some of the key steps involved in modelling flow behaviour may besummarized as follows:

    Identify the nature of the practical problem requiring solution; Identify the physical processes involved; Develop some mathematical representation of the physical processes; Understand any limitations to the mathematical representation;

  • 26 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Replace the (simplified) mathematical governing equations with a set of alternativeequations based, for example, on finite difference, finite element or finite volumeequations;

    Define the domain of the river to be studied and represent it by a suitable meshor grid. This process is sometimes known as the schematization;

    Define the boundary conditions to the domain, e.g. flows, velocities or water levelsat specified boundaries;

    Solve the set of equations using appropriate numerical techniques; Check the answers by solving any benchmark test cases to similar problems and

    compare the solutions; Calibrate the model for the particular problem in hand, optimizing key features

    such as water levels, velocity distributions, flow capacity, etc. by appropriateselection of any calibration coefficients;

    Check on the values and range of any calibration coefficients used, preferably bydirect measurement or by comparison with other values in similar circumstances,with reference to their physical meaning;

    Optimize the design or analysis by repeated use of the model.The word model is now used frequently in a broader sense, not just for the

    mathematical equations governing the physical processes, but to describe the wholemodelling process, as described above, often encapsulated in a particular softwarepackage. A good example of the development from physical processes to a concep-tual theoretical model is shown in the work of the ASCE Task Force on River WidthAdjustment (Thorne et al., 1998; Darby et al., 1998). Another example showing thedevelopment from initial data on overbank flow, through to conceptual thinking,model development and calibration issues is described by Knight (2008a&b).

    Modelling thus involves not only selecting the right tool for the right job, butalso being aware of the various steps involved, particularly concerning the limitationsand uncertainties in the chosen methodology. It is quite possible that, for example,the numerical procedures may vary between models even of the same type. Further-more, most models use empirical coefcients that have to be selected carefully at thecalibration stage, often without adequate field measurements on the particular riverin question. Turbulent parameters are also often based on inadequate full-scale mea-surements, sometimes even on flows other than in rivers, and in some cases on flowsnot even amenable to experimental scrutiny at all. See Nezu & Nakagawa (1993) andIkeda & McEwan (2009) for further details.

    The use of resistance coefficients to account for roughness effects is particularlyproblematic, as indicated by Yen (1991) and Morvan et al. (2008) and described laterin more detail in Section 2.2.3 and in Chapter 3. The engineer is thus faced with someconsiderable difficulty in blending theoretical and empirical knowledge together in apragmatic manner in order to simulate river processes and develop any type of modelat all. See Ikeda & McEwan (2009), Knight (1996 & 2008a&b) and Wang (2005) forfurther details.

    2.2.2 Schematisation of channel geometry

    The schematisation of river geometry may be relatively straightforward where thecross section is of a simple shape and prismatic. Figure 2.5 illustrates a trapezoidal

  • Practical and theoretical issues in channel hydraulics 27

    Figure 2.5 River Main, with compound trapezoidal cross section (i.e. a prismatic channel).

    Figure 2.6 River Severn in flood at Upton upon Severn, February 1990 (See colour plate section).

    compound section, an engineered channel, which forms part of a specially constructedflood relief scheme for the River Main in Northern Ireland. Between 1982 and 1986,an 800 m reach of the channel was reconstructed and realigned to form a doubletrapezoidal channel from Lisnafillan Weir to the junction with the Braid River. Thecross-section comprises a main channel with a bankfull depth of 0.91.0 m, a topwidth of 14 m, two berms covered with heavy weed growth, making a total width,inclusive of floodplains, of 27.330.4 m. The floodplains slope towards the mainchannel with a 1:25 gradient. The reach-averaged longitudinal bed slope is 0.00297.The river bed consists of coarse gravel, with a d50 of the order 1020 mm and the mainchannel side slopes consist of quarry stone (0.5 tonne weight, 100200 mm size). Thisreach has been closely monitored (Myers & Lyness, 1989) and subsequently modelledextensively, as discussed later.

    Many natural rivers, particularly in times of flood, exhibit much more variabil-ity in both cross section shape and plan form geometry than engineered channels.Figure 2.6 shows the lower reaches of the River Severn during extensive ooding inFebruary 1990, with the town of Upton upon Severn at the centre of inundation.In this case the river has overowed its banks and occupied part of its oodplain.Figure 2.6 thus illustrates the need to include sufcient oodplain area, together with

  • 28 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.7 River Severn in flood atWorcester, February 1990. Note land use management (See colourplate section).

    embankments, bridges and urban areas in any numerical model for effective flood riskmapping and management. Figure 2.7 shows an example of why it is important toinclude flood plain storage elements into any model. The flooded racecourse in theforeground will make a positive contribution towards reducing ooding, as it pro-vides additional short-term storage whenever particular floods occur. This illustratesa good example of integrated policy on land use. The reclaimed land in the centre, onthe other hand, may indicate poor land use policy, since it demonstrably restricts theflood plain width and hence exacerbates ooding locally.

    The severe economic damage that such flooding causes to infrastructure, the lossof life and the potential impact of climate change, have forced many politicians tore-consider land use policy. The global scale of the impact of flooding is described byKnight (2006a) and the impact of climate change is dealt with by Bronstert (2006a&b),Green et al. (2009) and Oshikawa et al. (2008). Examples of the impact of recentflooding in 7 European countries are described by Knight & Samuels (2007). TheEuropean Parliament ratified the Flooding Directive in 2007 (European Commission,2007) that requires the following for every river basin district:

    Preliminary flood risk assessment to identify areas for subsequent investigation; Flood risk maps; Flood risk management plans.

    The role of numerical models is therefore likely to play a signicant part in allof these activities. As described elsewhere by Knight (2006c), the schematisation ofa river, including its oodplain, is an important element in constructing a successfulmathematical simulation model. It should not be treated merely as an exercise in digi-tising numerous survey data, but rather as an art: that of blending the geometry andthe hydraulic flow features together in parallel. General rules for the location of cross-sections and the data requirements for 1-Dmodels are given in Samuels (1990&1995),Cunge et al. (1980), Casellarin et al. (2009) and Defalque et al. (1993), summarizedbelow:

    At model limits (boundary conditions) Either side of structures (for afflux or energy loss calculations)

  • Practical and theoretical issues in channel hydraulics 29

    At all flow and level measuring stations (for calibration purposes) At all sites of prime interest to the client Representative of the channel geometry About 20 B apart (as a first estimate only), where B = bankfull width A maximum of 0.2H/So apart A maximum of L/30 apart where L is the length scale of the physically important

    wave (flood or tide) The area for successive sections lies between 2/3 and 3/2 of the previous section. The conveyance for successive section lies between 4/5 and 5/4 of the previous

    sections.

    The overall plan form geometry of a river, and its associated oodplain bound-aries, as well as other large topographical features, will inevitably steer the flow intocertain patterns and produce large scale or macro flow structures. For example, densebankside vegetation or clumps of trees/hedges on the oodplain will create plan formvorticity that will govern lateral momentum transfers and mixing processes. The vari-ation in the shape of the main river channel around a meander bend likewise causessignicant variation in the longitudinal flow pattern. Convective accelerations anddecelerations along the channel will not only affect turbulence levels and resistance,but the balance of terms in the 1-D equation also. Likewise, non-prismatic oodplainswill produce transverse fluxes of mass and momentum which need to be considered(see for example Bousmar, 2002; Bousmar & Zech, 1999; Bousmar et al., 2004;Chlebek & Knight, 2008; Chlebek, 2009; Elliott & Sellin, 1990; Rezaei, 2006; Sellin,1995). The degree of sinuosity, as well as the depth of flow and the aspect ratio of theriver channel, will affect the position of the lament of maximum velocity, and thusinuence where bank erosion might take place (Okada & Fukuoka, 2002 & 2003).Such knowledge is clearly required before designing bank protection works, targetingmaintenance work or estimating flood wave speed (Tang et al., 2001).

    Although survey data are collected routinely at particular cross-sections, itshould be remembered that the hydraulic equations are generally applied to a reach(Laurensen, 1985; Samuels, 1990) and that energy losses are quantied over a specicdistance. Although care may be taken in identifying reaches that are approximatelyprismatic, there will inevitably be some longitudinal variation in shape and energygradient that will cause changes in hydraulic parameters in the streamwise direction.An example of the use of 1-Dmodels to determine the appropriate bankfull dischargealong a series of river reaches is given by Navratil et al. (2004).

    2.2.3 Roughness and resistance

    Although it is relatively straightforward to obtain digital terrain data, it is much harderto obtain the corresponding roughness data to go with the terrain data to the samelevel of detail (Defra/EA, 2003b; Hicks & Mason, 1998). For this reason the CESincludes a Roughness Advisor that allows the modeller to select appropriate rough-ness values for various substrates, types of vegetation and seasonal growth patterns.Figures 2.1(a) & (b) have already shown the variation that can occur in lowlandand upland rivers. To illustrate seasonal growth of instream and bankside vegetation,

  • 30 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figures 2.8 & 2.9 show two reaches of the River Blackwater in Hampshire, near Cole-ford bridge. The effect of this vegetation on the roughness coefcient and conveyancecapacity of the river throughout the year is documented by Sellin& van Beesten (2004).

    A distinction should be made between roughness, which denotes a surface textu-ral property and resistance, which is the cumulative effect of the roughness resistingany hydraulic flow over or through a particular type of roughness. The meaning ofthe term roughness in the field of uvial hydraulics, and how it is often formulatedas a resistance to flow term in 1-D, 2-D & 3-D numerical models, is described else-where more fully by Morvan et al. (2008). The Manning formula (1891), which wasintroduced in Section 2.1, is but one of a number of equations that attempt to quan-tify hydraulic resistance to flow. Although it is well founded and popular with riverengineers (Yen, 1991), it should be recognized that there are a number of alternativeformulations that describe resistance. A brief review of these and related topics is nowgiven.

    Figure 2.8 River Blackwater, showing seasonal growth (Winter to Summer) in a narrow reach (Seecolour plate section).

    Figure 2.9 River Blackwater, showing seasonal growth in a meandering reach (See colour platesection).

  • Practical and theoretical issues in channel hydraulics 31

    One of the earliest resistance equations was that by Chezy (1768), who expressedthe discharge in an open channel in terms of a roughness constant, C, where

    Q = CAR1/2S1/2o (2.23)This was followed by the Darcy-Weisbach equation (1857), which gave the head

    loss, hf , over a length, l, in either a circular pipe or open channel as

    hf = f U2

    2gd= f U

    2

    8gR(2.24)

    where hf = head loss, f = friction factor, l = length of pipe or channel, d = pipediameter, R = hydraulic radius (= A/P = d/4 for pipe running full), U = section meanvelocity (= Q/A), and Sf = friction slope (= hf /l). When the friction slope, Sf , is equalto the bed slope, So, (see Section 2.2.4 for further details), then Eq. (2.24) may beexpressed in a similar way to Eq. (2.23) as

    Q =(8gf

    )1/2AR1/2S1/2o (2.25)

    The Colebrook-White equation (Colebrook, 1939; Colebrook and White, 1937)clarifies how f varies in smooth and rough pipes, conduits running part full and openchannels. The variation of f is usually expressed in terms of two parameters, the relativeroughness, ks/d and the Reynolds Number, Re (= 4UR/). For clarity, two forms ofthe Colebrook-White equation are given, one for pipes and another for channels, butthese are essentially the same, since d = 4R:

    1f

    = 2.0 log[

    ks3.71d

    + 2.51(ud/

    )f

    ]

    (pipes & culverts running full) (2.26)

    1f

    = 2.0 log10[

    ks14.8R

    + 2.51(4UR/)

    f

    ]

    (open channels and pipes part full)

    (2.27)

    The variation of f with ks/d and Re is often shown plotted in the form of theso-called Moody diagram (1944), as illustrated in Figure 2.10. The roughness of anysurface is now characterised by ks, the so-called Nikuradse equivalent sand roughnesssize (see Nikuradse, 1933), a measure of the size of excresences, k, on a flat surfacethat would yield the same resistance as that in a circular pipe roughened with uniformgrains of sand. The Colebrook-White equation is physically well founded, since it tendstowards two theoretically limiting cases, described elsewhere by Schlichting (1979).For example, for very smooth surfaces, as ks 0, then Eq. (2.27) becomes the Prandtlsmooth law, in which f depends solely on Reynolds Number, Re, giving:

    1f

    = 2.0 log10(Ref ) 0.80 (2.28)

    On the other hand, as Re , Eq. (2.27) becomes the fully rough law, inwhich f is independent of Re and depends solely on the ratio of surface roughness, ks,

  • 32 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.10 Variation of friction factor, f , versus Re and ks/d or ks/4R (Moody, 1944).

    to hydraulic radius, R, i.e. ks/4R, giving:

    1f

    = 2.0 log10[

    ks14.8R

    ](2.29)

    This is particularly important in river engineering as it may be shown that thislinks Mannings n in Eq. (2.1) with ks, in Eq. (2.27), by the relationship:

    n = k1/6s[

    (R/ks)1/6

    18 log(11R/ks)

    ]

    (2.30)

    The Colebrook-White equation essentially expresses the transition region fromhydraulically smooth to hydraulically rough flow. Technically these terms aredefined by the shear Reynolds Number,

    (Uks/

    ), as

    Hydraulically smooth(Uks/

    )< 5

    Transitional flow 5 70

    where U is known as the shear velocity, defined by

    U =

    o/ (2.31)

  • Practical and theoretical issues in channel hydraulics 33

    with o = boundary shear stress and = fluid density. The shear stress and shearvelocity are important parameters in open channel flow as they frequently used whendealingwith sediment transport and erosion issues (Chang, 1988). The boundary shearstress may be shown to be related to geometric parameters as follows.

    Consider a short length of a prismatic channel, x in length, with a cross-sectionalarea, A, then in uniform flow, there will be a balance between the weight force actingdown the channel bed slope (= mgsin , where m = mass, = channel bed slope andtan = So), with the resisting force acting on the wetted area (=oPx, where o =mean boundary shear stress, and P = wetted perimeter). For small channel bed slopes(< 0.1), since sin tan = So, then for uniform flow gAxSo = oPx,noting that R = A/P, the mean boundary shear stress is given by

    o = gRSo (2.32)

    This is essentially the momentum principle applied to this particular case. Forpractical purposes, and in order to provide a link between shear stress and velocity (ordischarge), the energy principle needs to be invoked. If the shear stress, o, is assumedto be proportional to the square of the velocity (for dimensional reasons), then thissuggests that

    o U2 (2.33)

    The constant of proportionality must then be f /8, where f is the friction factor, inorder to satisfy Eq. (2.24). Hence

    o = f8U2 (2.34)

    It may then be observed that by equating Eqs. (2.31)& (2.34), theDarcy-Weisbachfriction factor, f , is simply the ratio of two velocities, U and U, since

    f = 8(UU

    )2(2.35)

    2.2.4 Energy and friction slopes

    So far, two longitudinal slopes have been introduced, the bed slope, So, and the frictionslope, Sf (= hf /l) via Eqs. (2.1) or (2.23) and (2.24) respectively. These only have thesame value when the depth and velocity do not change with distance, x, along thechannel, i.e. uniform flow exists and h = hn. Although this is an important typeof flow, often used when calculating conveyance capacity, as shown in the workedExamples 13, natural river flows are frequently affected by upstream or downstreamcontrols. This makes the flow non-uniform, that is to say, the depth and velocity nowvary with distance x along the channel, as well as possibly with time, t, if the flow isalso unsteady. See Figure 2.11 for general non-uniform flow behaviour and notation.

  • 34 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    x

    z

    Velocity head, U2/2g

    Velocity, U

    Friction slope, Sf

    Bed level

    Energy line

    Horizontal

    Water surface

    Datum level

    Bed slope, S0

    h

    z

    Figure 2.11 General non-uniform flow behaviour, with notation.

    Under these circumstances, the friction slope is given by the following 1-D equation

    Sf = So [hx

    + Ug

    Ux

    + 1gUt

    ](2.36)

    inwhich {h&U} vary as functions of distance and time {x& t}. Since the first two termson the right hand side are equivalent to the water surface slope, Sw, (=/x), thenEq. (2.36) may be simplified and expressed in a form useful for field measurements, as

    Sf = x

    Ug

    Ux

    1gUt

    (2.37)

    Equation (2.36) is important in extending rating curves (h vQ relationships), sinceit may be combined with Mannings equation, in which Sf replaces So in Eq. (2.1) fornon-uniform flow. The discharge in unsteady non-uniform flow, Q, in terms of thedischarge for steady flow at the same depth, based on the same Manning roughnesscoefficient, Qn, is then given by

    Q = Qn[1 1

    So

    hx

    UgSo

    Ux

    1gSo

    Ut

    ]1/2(2.38)

    The concept of specific energy that was introduced via Eq. (2.12) relates the depthand kinetic energy of the flow relative to the bed (invert) of the channel. If the levelof the channel bed (invert) above a horizontal reference frame is given by zb, then the

  • Practical and theoretical issues in channel hydraulics 35

    total energy, Et relative to this reference frame is

    Et = zb + h + U2

    2g= zb + E (2.39)

    The gradient of total energy in the downstream direction, Se, is then given by

    (Et)x

    = (zb + E)x

    = Se (2.40)

    in which the bed slope So = zb/x . In steady uniform flow, all three slopes,So, Sf & Se have the same value, S, and normal flow occurs (h = hn). Under thesecircumstances, the conveyance of a channel, K, is defined as

    Q = KS1/2 (2.41)Conveyance is thus a quantitative measure of the discharge capacity of a water-

    course, relating the total discharge, Q (m3s1) to a measure of the gradient or slopeof the channel. By comparing Eqs. (2.1) & (2.41), it may be seen that K (m3s1) isessentially related to the geometry and roughness of the channel. For example, whenMannings equation is used

    K = A5/3/(nP2/3) (2.42)For steady non-uniform flow, Eq. (2.36)may be reduced to an ordinary differential

    equation

    dhdx

    = So Sf1 Fr2 (2.43)

    Equation (2.43) may be solved to give the variation of depth, h, with distance, x,along a channel. The local depth is seen to depend on the longitudinal variation of 3parameters, the bed slope, So, the friction slope, Sf and the Froude Number, Fr. For areach of length, x, then the average friction slope, Sf , is required (Laurensen, 1985),and Eq. (2.43) is commonly written in the form

    x = ESo Sf

    (2.44)

    Equation (2.43) is usually known as the gradually varied flow (GVF) equation,and leads to a way of classifying longitudinal water surface profiles in open channels.For a wide rectangular channel, such that b >> h, then R h and Eq. (2.43) may bewritten as

    dhdx

    = So[1 (hn/h

    )10/3]1 (hc/h

    )3 (2.45)

    This shows that water surface proles depend upon values of So, hn and hc. Tra-ditionally, these proles have been divided into 5 classes, depending on the bed slope,

  • 36 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    and each class is further sub-divided depending on the depth of water, h, relative to thenormal and critical depths, hn and hc. For a given shape of cross section, discharge, bedslope and roughness, the values of hn and hc may be determined, as shown previously,and these will effectively divide the flow space into 3 depth zones. The water surface,being disturbed from its equilibrium normal depth, may lie in any of these 3 depthzones. Flow proles are then classied by denoting the type of channel bed slope by aletter and the depth zone by a number. For example, an M3 prole would imply thatthe channel bed slope is mild (M) and that the flow was in zone 3, i.e. supercritical.Since there are 5 classes of channel slope: horizontal (H), mild (M), critical (C), steep(S) and adverse (A), one might expect 15 possible proles. In fact 2 are inadmissible(H1 and A1) and C2 is technically uniform flow with hn = hc, so there are only 12distinct gradually varied flow proles.

    Of most interest to river engineers are those on mild or horizontal slopes, illus-trated in Figures 2.12 & 2.13 respectively. In Figure 2.12, the left hand side indicatesthe 3 depth zones, with schematic M1, M2 and M3 water surface proles sketched inrelation to imaginary hn and hc lines drawn parallel to the bed. On the right hand sideof the Figure, the smaller diagrams (a) to (f) represent some examples of where suchproles might typically occur in practice. For mild slopes (hn > hc), the backwaterprole (M1) is frequently encountered, particularly in relation to bridges and controlstructures, as shown later in Section 2.6. An estimate of the length of a backwaterprole has been given by Samuels (1989b), as

    L = 0.7h/So (2.46)

    where L = the distance upstream over which the water depths are disturbed by morethan 10% from their equilibrium condition (h = hn). The actual water surface profilemay be determined by numerical solution of the GVF equation, using the standardstep method for non-prismatic channels, or the direct step method for prismaticchannels. See Chow (1959) and Henderson (1966) for further details of these so called

    Figure 2.12 Gradually varied flow profiles on a mild slope.

  • Practical and theoretical issues in channel hydraulics 37

    Figure 2.13 Gradually varied flow profiles on a horizontal slope.

    step methods. Since there is a widespread need to calculate backwater lengths orprofiles, a backwater routine is included within the CES and further details are givenin Chapters 3 & 4 (Sections 3.2.4 & 4.3 respectively).

    In addition to the bed slope and friction slope, the energy slope, Se, is also impor-tant. Since this involves the kinetic energy correction coefcient, to account for the 3-Dnature ofmost flow fields in rivers, the next section considers a number of issues relatedto the velocity distributions within channels and their implications for 1-D modelling.

    2.2.5 Velocity distributions and implicationsfor 1-D modelling

    Up until now, the flow has been considered as one-dimensional (1-D), with the sectionmean velocity, U (= Q/A) being the only one used in all resistance equations andwater surface profile relationships. As mentioned earlier, this has certain implicationswith respect to energy, kinetic head, and approximating 3-D effects into a lateraldistribution model such as is used in the CES. Figure 2.14 shows a typical velocityfield for inbank flow in a simple trapezoidal channel, taken from the work of Yuen(1989). Figure 2.15 shows how the flow field becomes even more complex whenoverbank flow occurs, partly due to the strong interaction that may develop betweenthe flows in the main river channel and on the floodplain(s). Since 3-D concepts andflow structures are described in more detail in Section 3.1, only those that are relevantto the velocity field are introduced now, hopefully enabling the reader to understandmore about modelling flows in rivers.

    The first issue to appreciate is the 3-D nature of the flow, even in straight channels.Figure 2.14 shows a typical isovel pattern, secondary flow cells (Einstein & Li, 1958;Nezu & Nakagawa, 1993) and the resulting distribution of boundary shear stressaround the wetted perimeter (Knight et al., 1994). The maximum velocity may be atthe free surface for a wide river, as shown, but very often is just below the surface, par-ticularly in channels with a low aspect ratio (b/h). Boundary layers (Schlichting, 1979)develop from the bed and sides of the channel, with the local velocity, u, being zeroat every solid/liquid interface around the wetted perimeter (i.e. a boundary condition

  • 38 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.14 Isovels and boundary shear stresses for inbank flow in a trapezoidal channel (after Knightet al, 1994).

    Figure 2.15 Flow structures in a straight two-stage channel (after Knight & Shiono, 1996).

    of u = 0). If the channel is wide (large aspect ratio, b/h), then the velocity distribu-tion over a vertical at the centre-line position may be assumed to be approximatelylogarithmic, described by Prandtls logarithmic velocity law for a rough surface as

    uu

    = 1ln(

    zz0

    )(2.47)

  • Practical and theoretical issues in channel hydraulics 39

    where u = velocity a distance z above the bed, u is the local shear velocity, is vonKarmans constant, z0 = ks, = constant and ks is the Nikuradse equivalent sandroughness size. Thus z0 is assumed to be proportional to the size of the rougheningexcrescences, ks, as for smooth surfaces. Early experiments gave values of = 1/30and = 0.40 (now taken as 0.41). Hence the velocity distribution law for roughsurfaces is often assumed to be governed by an equation of the form

    uu

    = 2.5 ln(

    zks

    )+ 8.5 or u

    u= 5.75 log10

    (zks

    )+ 8.5 (2.48)

    The corresponding equation for velocity distribution over a smooth surface is ofthe general form

    uu

    = A ln(

    zks

    )+ B (2.49)

    where A & B are constants. Experiments show that B is a function of (uks/),given by

    B = 2.5 log10(uks

    )+ 5.5 (only valid for (uks/) < 5.0) (2.50)

    Combining Eqs. (2.49) & (2.50), and using Prandtls original constants, gives thecorresponding velocity distribution over a smooth surface as

    uu

    = 2.5 ln(uz

    )+ 5.5 or u

    u= 5.75 log10

    (uz

    )+ 5.5 (2.51)

    The second issue is to consider the implications of using a depth-averaged model.The depth-averaged velocity, Ud, is defined as

    Ud = 1h h

    0udz (2.52)

    Typical lateral distributions of Ud across a channel are shown in Figure 2.16. Asmight be expected, Ud = 0 at the two edges and increases towards the central region.The presence of secondary flows may mean that the maximum value of Ud does notactually occur at the centerline of the channel.

    The third issue is to consider how a 3-D velocity field might be included in a 1-Dformulation of energy or momentum. As noted earlier in Eqs. (2.12) & (2.17), this isusually achieved by use of the kinetic energy and momentum correction coefficients, and , here formally defined as

    = 1U3A

    Au3dA (2.53)

    = 1U2A

    Au2dA (2.54)

    These coefficients are introduced to ensure that the energy flux (=1/2 AU3)or momentum flux (=AU2) based on the area, A, and section mean velocity, U,

  • 40 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.16 Variation of Ud with lateral distance across a trapezoidal channel (Knight et al., 1994).

    Table 2.1 Typical values of kinetic energy and momentum correction coefficients.

    Correction Using the seventh powercoefficient Prismatic channels Rivers law, Eq. (2.55)

    1.15 1.30 (and higher) 1.045 1.05 1.10 (and higher) 1.016

    agree with the actual energy and momentum fluxes (= A 12u3dA and =A u

    2dA)obtained by summing over the whole area, using local velocities, u, applied to eachindividual element of area, dA. This effectively allows one to use the section meanvelocity, U, rather than knowledge of the complete velocity field (u = f {y, z}, whichcan be obtained from measurement or 3-D computational fluid dynamics (CFD) andintegrating over the areal surface. The use of these two correction coefficients has greatpractical merit in that only section mean velocities are used, and these are generallyknown. Typical values of and are shown in Table 2.1.

    As indicated in Table 2.1, a seventh power law is sometimes used to represent thedistribution of velocity over a vertical in turbulent open channel flow, rather than themore complex logarithmic laws, expressed earlier as Eqs. (2.48) & (2.51), on account

  • Practical and theoretical issues in channel hydraulics 41

    of its simplicity and closeness of fit to logarithmic distributions. This distribution maybe written as

    uumax

    =( zh

    )1/7(2.55)

    When this is inserted into Eq. (2.52), to obtain the depth-averaged value, it gives

    Ud = 78umax (2.56)

    Although Eq. (2.55) has the advantage that it is a simpler relationship to use thanEqs. (2.48) & (2.51), it has the disadvantage from an engineering perspective that itneeds the maximum velocity, umax to be known, as well as its position within the crosssection. If it is assumed that umax occurs at the free surface, then it may be shown thatUd occurs at approximately z = 0.4 h, regardless of which equation, Eq. (2.48), (2.51)or (2.55), is used. This is helpful, since it forms the basis of river gauging in openchannels.

    The fourth issue following on from the third, is the application of these and coefficients to flows in natural channels. As already indicated in Eq. (2.12), thekinetic head term based on the section mean velocity, U2/2g, needs to be multiplied inevery energy relationship by . In a similar way, wherever the momentum equation isused, the term QU should be multiplied by , as indicated in Eq. (2.18), to give thecorrect force-momentum relationship. The energy line (or surface) may be conceivedas an imaginary line drawn a distance U2/2g, above the water surface, as alreadyindicated by the dotted line in Figure 2.4. This is now shownmore clearly in Figure 2.17for flow in a section of open channel in which there is a constriction, such as arisingfrom bridge piers or a measuring flume. Upstream of the bridge, the velocities are

    Figure 2.17 Illustration of variation of kinetic heads and energy losses through a bridge (after Hamill,1999).

  • 42 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    lower and so the energy line is closer to the water surface than downstream wherethe velocities are larger. Note that the afflux is the increase in water level upstreamof a bridge, above the water level that would occur in the absence of the bridge, andshould not be confused with the energy loss across the whole bridge structure.

    In Figure 2.17 water surface dips as it passes through the section with a reducedwidth, on account of the increase in velocity from considerations of continuity (Q =U1A1 = U2A2). Since the frictional losses are proportional to the square of the velocity,the frictional losses in the narrow section will be larger than those in the channelupstream. Because of the change in cross section, there may also be additional energylosses in the contraction and expansion regions, with the losses in any expansionregion generally being larger than those in a contracting region. The overall losses arethen the sum of the frictional and additional energy losses, with the overall energyline indicating the total budget of energy and how it is degraded. A distinction shouldtherefore be made between the energy and friction gradients, Se and Sf .

    2.2.6 Hydraulic structures and controls

    A particular difculty in modelling water surface proles in rivers arises from the widevariety of hydraulic structures that are commonly used in river engineering. Theserange from weirs and flumes (Ackers et al., 1978; Bos et al., 1984), to bridges andculverts (Neill, 1973; Ramsbottom et al., 1997), gates and sluices (Franke & Valetin,1969; Kolkman, 1994; Larock, 1969; Rajaratnam & Subramanya, 1969; Rajarat-nam & Humpries, 1982), siphon spillways (Novak et al., 1996) and free overfalls(Chow, 1959). Each structure will cause a change in water level across the structureand should be modelled accordingly. For bridges, the head loss may be different fromthe afflux, as considered further in Section 2.6.

    Wherever the flow changes from subcritical to supercritical there exists a uniquerelationship between h & Q, and this may form one or two control points. Forexample, in the case of the sluice gate, illustrated in Figures 2.3 & 2.4 (Example 4 inSection 2.1), the control point on the upstream side of the gate controls all the waterlevels in the upstream direction, i.e. in the subcritical part of the flow. This is referredto as a backwater profile and, if the sluice gate were sited in a mildly sloping channel,

    Figure 2.18 Bridge arches in flooding River Avon (courtesy University of Birmingham) (See colourplate section).

  • Practical and theoretical issues in channel hydraulics 43

    Figure 2.19 Scour in channel with bridge (courtesy HR Wallingford) (See colour plate section).

    Figure 2.20 Flow downstream of the Holme sluices, River Trent (courtesy University of Birmingham)(See colour plate section).

    it would be an M1 profile (see Figures 2.12(a) & (c)). On the downstream side ofthe gate, at the vena contracta, the control point controls all the water levels in thedownstream direction, i.e. in the supercritical part of the flow, creating the H3 profilein Example 4.1 or an M3 as shown in Figure 2.12(c) if the channel has a mild slope.Knowledge about where control points occur, or rules about where they form, area pre-requisite for deciding which of the 12 gradually varied flow profiles will occurand for modelling any water surface profiles in open channels. Some examples of flowthrough bridges and some typical control structures are shown in Figures 2.182.22.

  • 44 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.21 Model of air regulated siphon Figure (model scale 1:10) (See colour plate section).

    Figure 2.22 Siphon outlet, showing flow over energy reducing blocks (See colour plate section).

    2.2.7 Calibration data for river models

    One of the main difculties frequently experienced in the development of a mathe-matical model of a river is the lack of suitable field data for calibration, especiallythat related to channel roughness, water surface behaviour and turbulence parame-ters under extreme flow conditions. Despite hydrometric survey work being routinelycarried out in the UK, there are generally never enough gauging stations in the rightposition, or indeed any monitoring of other key hydraulic parameters at all. This hastwo main consequences: that of making the model less useful than it might have beenand, secondly, the model results will have larger uncertainties associated with themthan perhaps originally hoped for. This issue has been discussed by several authors(e.g. Anastasiadou-Partheniou & Samuels, 1998; Knight, 2008a&b; Seed et al., 1993;Vidal et al., 2007) many of whom stress the need for a proportion of any modellingbudget to be spent on acquiring appropriate data. Well focussed physical model stud-ies, such as those undertaken in the UK Flood Channel Facility (Knight & Sellin, 1987;website www.flowdata.bham.ac.uk), are sometimes the only source of data at suf-ciently comprehensive spatial and temporal scales to be used in the effective calibrationof numerical models of river flow. These, and other sources of data, will be exploredfurther in Section 2.7.

  • Practical and theoretical issues in channel hydraulics 45

    In addition to the hydraulic variables, there are manymeteorological, hydrologicaland hydro-geological factors that affect the discharge in a river and also inuence itsvariability. Such factors might be related to rainfall-runoff issues, precipitation andsnow melt issues, groundwater issues, the underlying geology of the river basin, thenumber of tributaries and the behaviour of individual sub-catchments. The Flood Esti-mation Handbook (FEH, 1995) and the revised version (ReFEH, 2007) contain muchof this information that should be consulted by all modellers requiring further details.

    2.3 FLOW IN SIMPLE ENGINEERED CHANNELS

    This section, and subsequent sections in Chapter 2, now explores briefly the morecomplex flow conditions that can occur in actual rivers and watercourses and howthey might be modelled successfully. A full explanation of the scientific basis behindthe CES-AES software is given in Chapter 3, and more examples illustrating furtherpractical issues are given in Chapter 4.

    The modelling strategy used within the CES is largely based on the Shiono &Knight Method (SKM) of analysis, which gives the lateral distribution of depth-averaged velocity,Ud across channels of any prismatic shape. It is applicable to inbankand overbank flows in straight channels, and may be extended to flows in meander-ing channels by adding an additional term. The method captures certain 3-D floweffects and embodies them into a simpler 1-D approach. A full explanation of themethodology is given in Chapter 3, and further details of the model may be found inShiono & Knight (1988 & 1991), Knight & Abril (1996), Abril & Knight (2004), McGahey (2006), Knight, Omran & Tang (2007), Knight & Tang (2008), Mc Gaheyet al. (2006 & 2009) and Tang & Knight (2009). A brief introduction to the SKMapproach is now given, using H = water depth to make the notation appropriate tothe CES method (see Figure 2.23).

    The equations which govern the behaviour of uids in motion are known asthe Reynolds Averaged Navier-Stokes (RANS) equations. See Schlichting (1979) andDrazin and Riley (2006) for further details on the RANS equations. The governingReynolds Averaged Navier-Stokes equation for the streamwise motion of a fluid ele-ment in an open channel, with a plane bed inclined in the streamwise direction, maybe combined with the continuity equation to give

    [

    y(UV) +

    z(UW)

    ]= gSo + yx

    y+ zx

    z(2.57)

    where {UVW} = velocity components in the {xyz} directions, x-streamwise parallelto the channel bed, y-lateral and z-normal to the bed, = fluid density, g = gravita-tional acceleration, So = channel bed slope, and { yx, zx} = Reynolds stresses in thestreamwise direction on planes perpendicular to the y and z directions respectively.Figure 2.23 indicates these stresses, the notation and some of the key terms used.A physical interpretation of Eq. (2.57) would be: Secondary flows (streamwise andplanform) = Weight force + Reynolds stresses (lateral + vertical).

  • 46 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Figure 2.23 Flow in a natural channel and the various terms used in SKM and the CES.

    Shiono and Knight (1991) obtained a depth-averaged velocity equation byintegrating (2.57) over the flow depth (H), to give

    gHSo f8U2d

    (1 + 1

    s2

    )1/2+

    y

    {

    H2(f8

    )1/2Ud

    Udy

    }

    = y

    [H ( UV)d](2.58)

    where f = Darcy-Weisbach friction factor; = dimensionless eddy viscosity; s = thechannel side slope of the banks (1:s, vertical: horizontal) andUd = the depth-averagedvelocity. Solving Eq. (2.58) yields Ud as a function of y, the lateral co-ordinate. Otherterms are defined by:

    Ud = 1HH

    0

    U dz; b =(f8

    ) U2d ; yx = yx

    Udy

    ; yx = UH (2.59)

    whereU = (b/)1/2 = shear velocity. Eq. (2.58) thus includes the effect of secondaryflows (streamwise vorticity about horizontal axes) and planform vorticity (vortic-ity about vertical axes). Both these types of vortex structure are important in manyopen channel problems, and particularly so in overbank flow, as shown schematicallyin Figure 2.15. Based on experimental results, the term (UV)d is assumed to varyapproximately linearly, and therefore of the form

  • Practical and theoretical issues in channel hydraulics 47

    y[H ( UV)d] = (2.60)

    where is the lateral gradient of the advective term. In essence the SKM methodallows values of the three calibration parameters (f , and ) to be assumed for eachpart of the flow and then Eq. (2.58) to be solved either numerically or analytically. Inorder to illustrate its use, some simple examples are now given, beginning with flowin straight channels.

    Since the two simplest types of cross section that commonly occur, or are used inmodelling, are rectangular and trapezoidal, these are used first to demonstrate themod-elling strategy. See Figure 2.24 for examples of natural channels, and Figures 2.252.29for examples of laboratory channels used in refined calibration studies (Chlebek &Knight, 2006; Chlebek, 2009; Knight et al., 2007; Liao & Knight, 2007a&b; Omranet al., 2008; Shari et al., 2009; Tominaga et al., 1989).

    Figure 2.24 Inbank flows in approximately rectangular and trapezoidal river channels.

    Velocity

    0

    1

    2

    3

    4

    5

    0 2 4 6 8 10 12

    y (m)

    Ud

    (m

    /s)

    Bed Shear stress

    0

    10

    20

    30

    40

    50

    0 2 4 6 8 10 12

    y (m)

    Tau

    (N

    /m2 )

    (a) Lateral distribution of depth-averagedvelocity, Ud

    (b) Lateral distribution of bed shear stress, b

    Figure 2.25 Velocity and bed shear stress results for flow in one half of a 20 m wide rectangularchannel (single panel results, with y = 0 at centreline).

  • 48 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    (a) Lateral distribution of depth-averagedvelocity, Ud

    (b) Lateral distribution of bed shear stress,

    Velocity

    0

    1

    2

    3

    4

    5

    0 5 10 15 20 25y (m)

    Ud

    (m

    /s)

    Shear stress

    0

    10

    20

    30

    40

    50

    60

    0 5 10 15 20 25y (m)

    Tau

    (N

    /m2 )

    b

    Figure 2.26 Velocity and bed shear stress results for flow in one half of a 40 m wide rectangularchannel (single panel results, with y = 0 at centreline).

    Velocity

    00.10.20.30.40.50.60.70.8

    0 0.02 0.04 0.06 0.08y (m)

    Ud

    (m

    /s)

    lamda = 0.011lamda = 0.030lamda = 0.005

    Figure 2.27 Effect of on lateral distribution of velocity in a rectangular channel.

    The cross section of a rectangular channel may be treated singly or sub-dividedinto any number of panels, each of equal depth, but with varying flow parameters(f , and ) set for each panel. A trapezoidal channel may likewise be sub-divided, butthis time the panels will be of different depths if they are on the sloping side region,and constant in the flat bed region. In this case, the least number of panels to modelthe whole channel would be 2, one for each region, as shown in Figure 2.30 later. Thechoice concerning the number of panels is important and may affect the results in anumber of ways. If the flow and cross section are symmetric, as maybe sometimes forcertain flows in straight channels with rectangular and trapezoidal cross sections, thenonly half the channel requires analysis.

  • Practical and theoretical issues in channel hydraulics 49

    Velocity (direct)

    0.00

    1.00

    2.00

    3.00

    4.00

    5.00

    6.00

    y (m)

    Ud

    (m

    /s)

    gamma = 0gamma = 2 and 2gamma = 2 and 0Gamma = +20 and -20gamma = -20 and +20

    15.010.05.00.05.010.015.0

    Figure 2.28 Simulated bend flow in a rectangular channel (variable ).

    Velocity (direct)

    0.00

    0.50

    1.00

    1.50

    2.00

    2.50

    3.00

    3.50

    4.00

    4.50

    5.00

    0.0 2.0 4.0 6.0 8.0 10.0 12.0

    Ud

    (m

    /s)

    b1 = 5.0 m, f2 = 0.01b1 = 9.0 m, f2 = 0.01b1 = 9.9 m, f2 = 0.01b1 = 9.9 m, f2 = 1.0

    Figure 2.29 Flow in a 20 m wide rectangular channel with variable panel widths (2 panels, with y = 0at centreline).

    The philosophy adopted within SKM is to choose the minimum number of pan-els, commensurate with obtaining the maximum useful output from the results. Anygreater accuracy is unwarranted since it should be remembered that the methoditself is only an approximate one. However, there are a number of principles that

  • 50 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    HPanel 2

    yz

    Centreline

    Panel 1

    Figure 2.30 Cross-section of trapezoidal channel, schematized with two panels.

    should be followed concerning panel selection that are described in detail in Chapter3. For the present purposes, the following examples are restricted to single or, at most,a small number of panels.

    2.3.1 Flows in rectangular channels

    Example 6

    Determine the discharge and velocity distribution in a 20 m wide rectangular channelthat flows at a depth of 5.0 m at a bed slope of 0.001, assuming a Darcy f of 0.02,and constant values for and of 0.07 and 0.15 respectively.

    Answer & comment

    Consider only half the channel so one panel has a width of 10 m. The results are asshown in Figure 2.25 at 0.5 m intervals for y = 19 m, then at 0.1 m intervals to 9.9 mand finally at closer intervals near the wall. It can be seen that Ud = f {y} for the onepanel, even though the three calibration parameters (f , and ) are constant. Thedistribution of Ud is physically realistic, in that Ud = maximum at the centreline(y = 0) and zero at the wall (y = 10 m). The method thus mimics the antici-pated shear layer, and also gives the bed shear stress distribution since, once Ud isknown, values of b may be calculated from Eq. (2.59). Both distributions are hyper-bolic, and this illustrates how some details of the flow have now been calculated,in a way that was not possible using the simple 1-D Manning resistance equation,Eq. (2.1).

    Having determined the lateral distribution of Ud, the individual values maybe multiplied by the appropriate area of each sub-panel and the discharge obtained.This gives Q = 330 m3s1, which is somewhat different from the value of 362 m3s1calculated below, on the basis of using the same value of f in Eq. (2.25).

  • Practical and theoretical issues in channel hydraulics 51

    From Eq. (2.25):

    U = (8gRSo/f )1/2 = (8 9.807 3.3333 0.001/0.02)1/2 = 3.616 ms1.Q = UA = 3.616 20 5 = 361.6 m3s1.

    The same value ofQ should not in fact be expected, despite Eq. (2.25) being essentiallya single channel method (SCM), since the influences of the other two parameters( and ) have not yet been considered. For example, halving the value of (i.e. lesslateral shear) and putting = 0 (no secondary flow) in Eqs (2.58) & (2.59) gives Q =364 m3s1 which is much closer to the value of 362 m3s1 calculated using Eq. (2.25).

    The lateral distribution of velocity in Figure 2.25(a) is seen to give Umax =4.0 ms1 at the centreline and a mean value of U 3.2 ms1. The mean value isagain slightly different from the value calculated from Eq. (2.25), U = 3.616 ms1,for the reasons given above. The centre-line value of bed shear stress in Figure 2.25(b)is seen to be approximately b 40.0 Nm2, which is lower than the mean bed valuebased on assuming two-dimensional flow, (b2d = gHS0) and higher than the meanboundary shear stress for the whole cross-section ( o = gRSo). This is to be expected,since:

    b2d = 1000 9.807 5.0 0.001 = 49.03 Nm2. o = 1000 9.807 3.3333 0.001 = 32.69 Nm2.

    The value of maximum bed shear stress will approach the two-dimensional value(b = gHS0) as the channel is widened. For example if the width is now doubled to40 m, and assuming the same values for the parameters (f , and ) as before, givesthe distributions shown in Figure 2.26. In this case, the centre-line bed shear stress,b, is now much closer to the two-dimensional value of 49.03 Nm2, since the aspectratio of the channel (width/depth) has risen to 40/5 = 8. This is still not technicallya wide channel as the simulation results here are some way from those for a widechannel (see Figure 9 of Knight & MacDonald (1979), Knight et al. (1992 & 1994)and Al-Hamid (1991) for further details). For practical purposes the aspect ratio needsto be at least >10 to get within 95% of b2d and generally >25.

    Figure 2.26(a) shows Ud = f {y} for the 40 m wide channel, varying fromUmax = 4.38 ms1 at centerline to zero at the wall. Integrating across the width givesQ = 774.9 m3s1. This is less than the value of Q = 792.24 m3s1 below, based onEq. (2.25):

    U = (8gRSo/f )1/2 = (8 9.807 4.0 0.001/0.02)1/2 = 3.9612 ms1.Q = UA = 3.9612 40 5 = 792.24 m3s1.

    Lowering to 0.035 causes the values of Ud to rise, especially nearer the walls, givingQ = 810.3 m3s1, now above the value of 792.3 m3s1. This example has shown thatthe individual values of the parameters and , as well as f , will also influence thedischarge, and so this is explored further in the next example.

  • 52 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    Example 7

    Using only one panel to represent half of a rectangular channel, explore the combinedinfluence of all 3 parameters (f , and ).

    Answer & comment

    In order to test the methodology further, data are needed on channels where bothvelocity and boundary shear stress distributions have been measured with knownaccuracy. Since such data are rare at full scale in large channels, recourse has to bemade to data from small scale laboratory channels where it had been possible tomeasure both parameters accurately. One set of data that is suitable is highlighted byChlebek & Knight (2006), taken from experiments in a series of small rectangularchannels. Further discussion on data and availability are given in Section 2.7.

    The general effect on Q of varying all 3 parameters and the percentage of shearforce on the channel walls, %SFw, is shown in Table 2.2, based on numerical simula-tions of all the experimental results from the various channels, covering a wide rangeof aspect ratios. The %SFw values were obtained by integrating the bed shear stressesover the semi width of the channel, determining the overall shear force on the bed,SFb, and then subtracting it from the overall mean shear force for the whole channel,(o P) to give SFw, which is then expressed as a percentage of the total (bed andboth walls).

    Figure 2.27 shows the effect of varying on the lateral distribution of Ud, whileholding f and constant for one particular simulation. In this case, the experimentalchannel had a width of 152 mm, a depth of 200 mm, a Q of 13.78 l/s with a valueof %SFw of 73.08%. The %SFw value is understandably high, given the narrownessof the channel (aspect ratio of only 0.76). In general, it was found that for a givendepth, when was increased, with fixed f & , it caused Q to decrease and %SFwto increase. The reverse occurred when was decreased. Tables 2.2 & 2.3 illustratethese changes quantitatively, based on this one depth (H = 200 mm) with referencevalues of Q = 13.78 ls1 and %SFw = 73.08% for which f = 0.018, = 0.011 and = 0.47 gave the closest approximation to the experimental data, using just one panel.

    The results in rows 2 & 3 of Table 2.3 indicate that the effect of varying f from0.040 to 0.005 (from optimum of 0.018) is to change Q from 13.78 ls1 to 10.46 and

    Table 2.2 Effect of varying one calibration parameter, holding the othertwo constant.

    Variableparameter Fixed parameters +/ Q %SFwf & increase decreases decreases

    & decrease increases increases f & increase decreases increases

    f & decrease increases decreases f & increase decreases increases

    f & decrease increases decreases

  • Practical and theoretical issues in channel hydraulics 53

    Table 2.3 Effect of varying parameters f , and on values of Q and %SFw(Highlighted row shows closest simulation to the experimentaldata, i.e. optimum values).

    Q %SFwf (litres sec1) (%)

    0.018 0.011 0.047 13.78 73.080.040 0.011 0.047 10.46 65.230.005 0.011 0.047 20.52 83.080.018 0.030 0.047 9.36 87.540.018 0.005 0.047 17.35 57.440.018 0.011 0.500 11.97 79.680.018 0.011 0.000 13.95 72.93

    20.52 ls1 respectively, and to change%SFw from 73.08% to 65.23% to 83.08%. Theeffect of varying from0.030 to 0.005 (fromoptimumvalue of 0.011) is shown in rows4&5, and the effect of varying from0.500 to 0.000 (fromoptimumvalue of 0.047) isshown in rows 6& 7. This example highlights the need to consider the combined effectof all 3 calibration parameters on two objective functions (discharge and a boundaryshear force, i.e. Q and %SFw). Before considering this in detail, and using substantivesets of data outlined in Section 2.7, the influence of the secondary flow term, , as wellas the appropriate number of panels to be used to represent the channel cross section,also need to be investigated. These are now considered together in Example 8.

    Example 8

    A rectangular channel is 20 m wide, 3.0 m deep and has a bed slope of 0.001. Usingtwo panels to represent half of this channel, investigate the influence of the secondaryflow term, , and panel selection on the flow results, assuming f = 0.01 and = 0.10for all panels.

    Answer & comment

    In the absence of suitable full scale data, particularly those in which has been mea-sured, this exercise on secondary flow effects and the number of panels that should beused has to be a purely numerical one. Detailed numerical simulation of multi-panelsis reserved for an example in Section 2.3.2, using laboratory data for both Ud and%SFw and examples in Chapter 3 using river data, for which data exists on Ud butnot on boundary shear stress distributions.

    Figure 2.28 shows the results of these simulations for a 20 m wide channel inwhich two panels of equal width (b = 10m) make up the total width of the rectangularchannel. This is equivalent to modelling half the channel with just a single panel 10 mwide. The results show that for = 0, the distribution ofUd is symmetric, as expected.For = 2, in all panels the velocities are diminished slightly. When is changedin sign for each panel either side of the centerline, i.e. 1 = +20 on one side and2 = 20 on the other side, then the velocity distribution is no longer symmetric, but

  • 54 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    biased towards one side or another. Such a distribution is that which might occur ina left-handed or right-handed bend, with a higher velocity on the outside of the bend.Not surprisingly this is seen to be produced here by altering the secondary flow term, .As shown later in Chapter 3, there are other ways of attempting to simulate flow inmeandering channels.

    Figure 2.29 shows the effect of splitting a single panel into a number of sub-panels.In this case, a single panel of width b, equivalent to half of the channel width, is dividedfurther into 2 sub panels, of widths b1 and b2, to make up half of a rectangularchannel of semi-width, b, thus making b2 = b b1. The width of one panel, b1, isaltered progressively from 5.0 m to 9.9 m in order to see if the smaller panel adjacentto the wall captures the very high shear in that region. As can be seen, with b1 =9.9 m and b2 = 0.1 m the results compare quite well with the simulation undertakenwith two equal sized panels. The effect of increasing the roughness in the smallerpanel adjacent to the wall by a factor of 100 is seen to reduce the velocity in thatregion, thus indicating a possible way of refining any calibration to fit data. This isdealt with more thoroughly in the next section that deals with flows in trapezoidalchannels.

    Example 9

    Attempt some of these exercises using the CES software, which is based on the SKM,and compare your answers with those given in this Chapter. All the results in Exam-ples 68 were obtained using spreadsheets to evaluate the analytical solutions toEq. (2.58) directly.

    2.3.2 Flows in trapezoidal channels

    The previous examples have indicated some general features of the SKM for flows inrectangular channels. The use of multi-panels is now explored more thoroughly, usingdata from laboratory experiments undertaken by Yuen (1989) and Tominaga et al.(1989). Figure 2.30 shows a typical schematization for a trapezoidal channel, basedinitially on just two panels. Figure 2.31 shows the results of a simulation based on usingsuch a schematization, together with some experimental data. Although Figure 2.31(a)shows that the lateral distribution of Ud is reasonably well simulated, Figure 2.31(b)shows that the corresponding lateral distribution of boundary shear stress, b, is notsimulated as well. A much better simulation for the latter is obtained when each of thepreviously used 2 panels are sub divided further, making 4 panels in all, as shown bythe results in Figures 2.32(a) & (b). It was found by Omran (2005) that 4 panels arenormally quite adequate for practical purposes when modelling most types of flow intrapezoidal channels.

    In particularly demanding cases, where even greater precision is required, such aswhen attempting to simulate the boundary shear stress distribution very accurately in

  • Practical and theoretical issues in channel hydraulics 55

    (a) Velocity distribution (b) Boundary shear stress distribution

    Depth-averaged velocity distribution

    0.00.10.20.30.40.50.60.7

    -0.25 -0.20 -0.15 -0.10 -0.05 0.00y (m)

    Ud

    (m

    /s)

    experimentalSKM

    Shear stress distribution

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    -0.25 -0.20 -0.15 -0.10 -0.05 0.00y (m)

    Tau

    (N

    /m2 )

    experimentalSKM

    Figure 2.31 Results of velocity and boundary shear stress simulations of flow in a trapezoidal channel(2 panels). [Exp.16, Yuen, 1989, with f1 = 0.0185, f2 = 0.020, 1 = 2 = 0.07, 1 = 0.5,2 = 0.2].

    (a) Velocity distribution (b) Boundary shear stress distribution

    Shear stress distribution

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    -0.25 -0.20 -0.15 -0.10 -0.05 0.00

    y (m)

    Tau

    (N

    /m2 )

    experimentalconstant f

    Depth-averaged velocity distribution

    0.00.10.20.30.40.50.60.7

    -0.25 -0.20 -0.15 -0.10 -0.05 0.00

    y (m)

    Ud (

    m/s

    )

    experimentalSKM

    Figure 2.32 Results of velocity and boundary shear stress simulations of flow in a trapezoidal channel(4 panels, Exp 16, Yuen), after Knight, Omran & Tang (2007).

    corner regions, then 6 panels may be required. Under these circumstances, the number,pattern and strength of the secondary flow cells should be taken into account, asshown in Figure 2.33 by Tominaga et al. (1989). The results of simulating the sameexperiment shown in Figures 2.31& 2.32 are shown in Figure 2.34. It can be seen thatby using 6 panels the sharp decrease in boundary shear stress in the corner region (y =0.075m) ismuch better simulated. Another feature that is importantwhen dealingwithboundary shear stress with multiple panels of different roughness is to include somelateral smoothing into the friction factor, f , in order to overcome any discontinuity inb that arises from use of a depth-averaged approach based on Eqs (2.34) & (2.52).This is shown in Figure 2.34, where lateral distributions of boundary shear stress, b,are shown for constant f values (saw tooth plot, 4 panels), linearly varying f values

  • 56 Practical Channel Hydraulics: Roughness, Conveyance and Afflux

    b/2b/2b/2 b/2

    ++ -+ --

    P1P2P2aP3a P3 P4

    +--++-

    Schematic variation of (UV)d

    Sign of y(UV)d

    Panel number

    Sign of +--++-

    Figure 2.33 Secondary flow cells in a trapezoidal channel, after Tominaga (1989) and Omran (2005).

    Shear stress distribution

    0.500.550.600.650.700.750.800.850.90

    -0.25 -0.20 -0.15 -0.10 -0.05 0.00y (m)

    Ta

    u (

    N/m

    2 )

    experimentallinear fmodified linear fconstant f

    Figure 2.34 Results of boundary shear stress simulations of flow in a trapezoidal channel (6 panels,with variable , Exp. 16, Yuen), after Omran (2005).

  • Practical and theoretical issues in channel hydraulics 57

    (4 panels) and modified linearly varying f values (6 panels). Further details of this andrelated issues may be found in Omran (2005), Omran & Knight (2006), Knight et al.(2007) and Omran et al. (2008). Figures 2.312.34 thus indicate that the main issuesthat need to be considered are the number of panels and the values of each parameterwithin individual panels.

    2.4 INBANK FLOW IN NATURAL RIVERS

    The previous section has shown some of the background to modelling flows usingSKM/CES approaches, applied to simple prismatic channels, and how useful somelaboratory data might be for validation purposes. However, river engineers are usuallyconcerned with flows in natural rivers, where the cross-section is generally irregular inshape, not necessarily prismatic, the distribution of roughness more varied around thewetted perimeter, and other practical issues such as vegetation, culverts and bridgesneed to be considered. This section and subsequent ones now address all of theseconcerns and illustrates them via several case studies.

    Figure 2.23 shows that the flow in a natural channel is more complex than thosecases considered in Section 2.3. Firstly the irregular cross section has to be schema-tised using survey data to create multiple panels, and should be undertaken with someknowledge of the flow characteristics. Secondly, because there is often a lack of knowl-edge concerning the velocity distributions from suitable field data, care needs to betaken over the choice of the 3 calibration parameters (f , and ) to be used in eachpanel. Arising from this, a measure of uncertainty in the predicted results also needs tobe established. These features are now illustrated using the CES applied to an actualriver, the River Ngunguru in New Zealand.

    The example of the River Ngunguru at Drugmores Rock has been deliberatelychosen to illustrate inbank flow in a natural channel that has many of the difficultfeatures cited above. It will also serve to highlight further modelling issues, and istotally unlike any of the previous examples, based on simple prismatic laboratorychannels. Furthermore, one particular reach has been subject to ongoingmeasurementssince 1969 (Hicks & Mason, 1998). The Ngunguru River drains a catchment area of12.5 km2, has a mean annual flood discharge of 61 m3s1 and an average flow rateof 0.41 m3s1. Water level and discharge measurements were taken at three cross-sections along an 80 m reach, with a gentle bend in the downstream portion. Thesections were approximately 12 m wide and 2 m deep. The water surface slope variedfrom 0.0037 at low depths to 0.0064 at large depths. The river bed consists of graveland cobbles and the banks are lined with grazed grass and scattered brush, as shownin Figure 2.35. The observed Manning n values vary with depth in the range 0.051to 0.160, with the values increasing substantially at lower depths. The cross section isshown in Figure 2.36.

    The cross section was divided into 100 panels or elements and values of the 3calibration parameters (f , and ) chosen for each panel, based on the rules outlinedin Chapter 3. The CES software was then used at a specified depth to solve for thelateral distribution Ud v y, which was subsequently integrated laterally to give thedischarge at that same value of depth. Repeated application over a range o