Shallow Water Hydraulics
Shallow Water Hydraulics
Shallow Water Hydraulics
Willi H. Hager
Shallow
Water
Hydraulics
Shallow Water Hydraulics
Oscar Castro-Orgaz Willi H. Hager
•
123
Oscar Castro-Orgaz Willi H. Hager
University of Córdoba VAW, ETH Zürich
Córdoba, Spain Zürich, Switzerland
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To our Families
Preface
The purpose of Shallow Water Hydraulics is to present the theory and the com-
putation of open channel flows using analytical, numerical, and experimental
results. The book’s target audience includes graduate and undergraduate students as
well as practitioners of civil and environmental engineering. The book is concep-
tually divided into four parts: fundamental equations (Chap. 1), steady open channel
flows (Chaps. 2–4), unsteady open channel flows (Chaps. 5–9), and special topics
(Chaps. 10 and 11). As supporting learning material for students, a library of
numerical codes is available in Chap. 12. The complexities of theory and numerical
methods are progressively increased toward the end of the book, rendering pieces
of the material useful for courses of different levels.
It appears impossible to start the presentation of a new open channel flow book
without acknowledging the former works of Ven Te Chow (Open Channel
Hydraulics, 1959) and Francis M. Henderson (Open Channel Flow, 1966). Despite
Chow produced an encyclopedic treatise, it is Henderson’s outstanding book that
the authors followed. His presentation of concepts is simply unique, and although it
was published more than 50 years ago, it is still a main reference. Credit deserves
also the books of Subhash C. Jain (Open Channel Flow, 2001) and Hubert Chanson
(The Hydraulics of Open Channel Flow: An introduction, 2004), similar to
Henderson’s, pursuing a clear and brilliant presentation of open channel flow
concepts. These books offer a first open channel flow course, covering topics from
steady flow to fundamental questions relating to unsteady flow. However, they are
lacking a detailed exposition of numerical methods to solve the unsteady open
channel flow equations. The books of Sergio Montes (Hydraulics of Open Channel
Flow, 1998), M. Hanif Chaudhry (Open Channel Flow, 2008), and Eleuterio Toro
(Shock-Capturing Methods for Free Surface Shallow Flows, 2001) cover these
aspects in detail and can be used in an advanced open channel flow course. The
above-quoted books were used by the first author to offer a course on Operation of
Rivers and Reservoirs in the Environmental Hydraulics Master program held at the
University of Córdoba, Spain. The purpose was to give students a complete
overview on the computation of open channel flows, starting with the basic
equations, continuing with the solution of steady open channel flows, penetrating
vii
viii Preface
into the computation of unsteady open channel flows using modern numerical
methods, and finally introducing advanced topics including flows on movable beds
and sediment transport and non-hydrostatic flow modeling. The material of this
book originated from the lecture notes prepared for this course, and deep recog-
nition is therefore given to the books quoted above, which were used as a source of
knowledge and inspiration for years.
The authors are of the opinion that it is better to start from generalized equations
and then simplify the results using suitable approximations. From this perspective,
the fundamental equations of open channel flow are settled producing a rigorous
vertical integration of the 3D RANS equations (Chap. 1). The resulting mathe-
matical statements are then simplified for selected flows; in particular, the impli-
cations of assuming the hydrostatic pressure distribution are discussed in detail. The
emphasis of the book is directed to one-dimensional problems. The corresponding
steady flows are therefore described using the energy and momentum principles,
and a variety of numerical solutions to the algebraic and ordinary differential
equations governing these flows are presented (Chaps. 2 and 3). The transitions
across the critical depth, along with the formation of singular points in weir flow
and the hydraulic jump beyond a sluice gate, are detailed (Chap. 4). The funda-
mental equations of unsteady open channel flows along with their continuous and
discontinuous solutions are then presented, and modern shock-capturing
finite-difference and finite volume methods to solve them are extensively descri-
bed (Chaps. 5–9). A detailed description of dam break waves (Chap. 6) and sluice
gate maneuvers in open channels treating them as the solution of a Riemann
problem (Chap. 8) is provided. Special topics selected are the inclusion of sediment
transport and movable beds in shallow water models, as used to predict geomorphic
dam break waves (Chap. 10), and the computation of steady and unsteady
non-hydrostatic free surface flows (Chap. 11).
A collection of hand-solved exercises along the book is not presented. Rather,
the book offers to students and instructors a collection of source codes where each
type of problem discussed in the book is implemented step by step. The collection
of source codes is written in Visual Basic, and each code is inserted as a macro in
Microsoft Excel®. Teaching experience indicated that this approach permits stu-
dents an easy use of the material, and a productive interaction with the lecturer
during the classes is generated.
The authors hope that after studying the book, the reader will have a solid
background on the theory and computation of steady and unsteady open channel
flows, permitting to advance to the study of more complex problems relating to
two-dimensional numerical modeling of free surface flows.
ix
Contents
xi
xii Contents
1.1 Introduction
Open channel flow is the study of the movement of liquids with a free surface,
which is by definition an interface in contact with the atmosphere. Beautiful
examples of these flows are seen at dam spillways (Fig. 1.1a) or along rivers
(Fig. 1.1b).
While these flows can be tackled with techniques of general fluid mechanics, this
book is devoted to the use of approximate methods widely employed in civil and
environmental engineering, namely the use of shallow water models. These refer to
depth- or cross-sectional averaged models where the variation of the field variables
in the vertical direction is only approximately accounted for, or even neglected
(Vreugdenhil 1994; Toro 2001). Typical open channel flow problems are solved by
assuming one-dimensional (1D) or two-dimensional (2D) flows in a horizontal
plane. Common open channel flow sections in practice are irregular as in a river
(Fig. 1.2a), trapezoidal for a hydropower canal (Fig. 1.2b), rectangular for a
spillway chute (Fig. 1.2c) and circular for a sanitary sewer (Fig. 1.3d), among
others.
To classify open channel flows, assume for the moment 1D flow. While the
water moves in a channel, there are variations in depth and velocity both in space
(coordinate x) and time t. Let the water depth be h and the velocity U without
rigorously defining these physical quantities at the current introductory stage. An
open channel flow is classified based on the following features:
Steady and unsteady flows
A flow is steady if the flow variables do not change with time at a given position. If
these change with time at a fixed position, the flow is referred to as unsteady. The
flow is said also to be variable as it changes with time.
Fig. 1.1 Examples of free surface flows a spillway of Aldeadávila Dam; photo taken from https://
es.wikipedia.org/wiki/Presa_de_Aldeadávila, b Amazon river flowing through the Amazon rain-
forest (photo of public domain by NASA)
Fig. 1.2 Typical channel flow sections a irregular, b trapezoidal, c rectangular, d circular
Fig. 1.3 Open channel flow examples a steady uniform, b steady gradually varied, c unsteady
uniform, d unsteady gradually varied, e unsteady rapidly varied, f varied flow over a dam and in
tailwater channel
mixture effects are neglected, the pressure distribution within the hydraulic jump
can be assumed to be hydrostatic (Khan and Steffler 1996a; Castro-Orgaz and
Hager 2009). The flow is therefore rapidly varied and hydrostatic. Beyond the
jump, the water surface profile in the nearly horizontal tailwater channel has a small
variation of depth with distance, resulting in negligible streamline curvature. The
flow is therefore gradually varied and hydrostatic. In the vicinity of the free drop,
1.1 Introduction 5
the variation of flow depth with distance is large, the streamline curvature is sig-
nificant and the pressure distribution non-hydrostatic. The flow is therefore rapidly
varied and non-hydrostatic.
The main purpose of this book is to present the theory and computation of open
channel flows using the hydrostatic pressure approach. This type of model can be
collectively named shallow water model. Both steady and unsteady flows will be
tackled, as well as sediment transport problems. Unsteady flow computations will
include both finite-difference and finite volume methods, focussing on 1D models.
Extension to non-hydrostatic flows is considered in Chap. 11.
In civil and environmental engineering applications, free surface flows are often
treated using the one-dimensional (1D) continuity, momentum, and energy equa-
tions, in which information on the field variables across the flow section are
averaged (Chow 1959; Henderson 1966; Liggett 1994; Montes 1998; Sturm 2001;
Jain 2001; Chaudhry 2008). de Saint-Venant (1871) and Boussinesq (1877) are the
fathers of depth-averaged open channel flow modeling, both proposing approximate
models (Castro-Orgaz and Hager 2017). The rational development of 1D flow
6 1 Fundamental Equations of Free Surface Flows
models starts with the 3D equations of continuity, momentum and energy, which
are integrated across a section normal to the channel bed (Keulegan 1942; Strelkoff
1969; Yen 1973). Prior to the averaging process, the full Navier–Stokes equations
are time-averaged for a turbulent flow, resulting in the Reynolds equations. This
development was first presented by Keulegan (1942) and Keulegan and Patterson
(1943). Further studies were conducted by Jaeger (1956), Chen and Chow (1968),
Yen (1975), Lai (1986), Liggett (1994), Montes (1998), Jain (2001) and Chaudhry
(2008). The contributions of Strelkoff (1969) and Yen (1973, 1975) are notable,
because they detail the fundamental differences between the 1D energy and
momentum equations, retaining the flow equations as a function of averaging
coefficients accounting for arbitrary distributions of velocity and pressure.
toward the basic laws of similitude involving density currents and the mixing
of salt with freshwaters. Keulegan investigated both questions thoroughly and
presented classic papers relating to the lock exchange experiments. His results
on wave propagation and density currents were published in the 1950 book
Engineering hydraulics of Rouse, and the 1966 book Estuary and coastline
hydrodynamics of Ippen, respectively. Keulegan was awarded a number of
prestigious decorations, including the National Medal of Science, Honorary
Membership of ASCE, and in 1979 election to the National Academy of
Engineering.
The coordinate system considered by Strelkoff (1969) and Yen (1973) to pro-
duce the cross-sectional averaged Reynolds equations was orthogonal and fitted to
the terrain. This system is widely used in geophysical flow research (Savage and
Hutter 1989, 1991; Pudasaini and Hutter 2007). Strelkoff (1969) and Yen (1973)
transformed the Navier–Stokes equations in Cartesian coordinates to the
terrain-fitted reference system stated by means of a rotation of axes to render two of
the coordinates tangent to the terrain, and the third normal to it. While it is feasible
to produce such a transformation from Cartesian to terrain-fitted coordinates, a
simple axes rotation is not the exact mathematical operator, so that the channel flow
equations developed by this procedure are not general. Dressler (1978) transformed
the plane Euler equations in Cartesian coordinates to a boundary-fitted system of
reference by using a Jacobian matrix. In his modified form of the Euler equations,
the channel bed curvature appears. This feature is also observable in the granular
mass flow equations by Savage and Hutter (1989, 1991). The discrepancy between
Dressler (1978) and Savage and Hutter (1989, 1991), and the form of the Navier–
Stokes equations to obtain 1D equations used by Strelkoff (1969) and Yen (1973),
indicate that the equations developed by the latter apply strictly to flows over
constant slope channels. The development by Strelkoff (1969) is also available from
Jain (2001). The key message is that the Navier–Stokes equations in the devel-
opments of Strelkoff (1969) and Yen (1973) should be expressed in the terrain-fitted
system of coordinates using a Jacobian matrix rather than a simple rotation of axes,
because only then the averaging process is to be conducted. If the developments of
Strelkoff (1969) and Yen (1973) are repeated with these considerations, additional
terms including the terrain curvature appear, as those contained in the
depth-averaged avalanche flow equations by Savage and Hutter (1989, 1991). The
transformation of the Reynolds equations to any curvilinear system (orthogonal or
not) is comprehensively reported by Rouse (1959) and Schlichting and Gersten
(2000). A disadvantage of the 1D channel flow equations of Strelkoff (1969) or Yen
(1973) is that undetermined averaging coefficients need to be mathematically closed
for computations. These are difficult to evaluate, like those of the 1D energy
equation for turbulent unsteady flow.
An alternative to terrain-fitted coordinates is the resort to Cartesian coordinates,
despite this option was less used in open channel flows. Notable contributions to be
8 1 Fundamental Equations of Free Surface Flows
considered below are the works by Steffler and Jin (1993), Khan and Steffler
(1996b, c), Denlinger and Iverson (2004), Iverson (2005), Denlinger and O’Connell
(2008), and Iverson and Ouyang (2015). Note that the generalized free surface flow
equations in Cartesian coordinates were given by Yen (1975), thereby removing the
previous problem by the use of terrain-fitted coordinates.
Due to the increase of mathematical complexities while doing an averaging
process in terrain-fitted coordinates, Cartesian coordinates will be generally used in
this book. Further, given the difficulties in using the energy equation for
depth-averaged turbulent unsteady flow computations over 3D terrain, the use of
this (important) principle is limited to steady flow problems, and therefore pre-
sented in Chap. 2. The general 3D flow equations of mass and momentum will be
depth-integrated to produce a general 2D averaged model in this chapter. This
general formulation applies in practice to produce 2D computational models
making suitable approximations to the field variables. For the case of 1D flow, the
2D-averaged flow equations will be laterally integrated, producing the corre-
sponding cross-sectional averaged model. Sediment transport and movable beds are
finally considered.
Here the general depth-integrated equations for a mixture of fluid and sediments,
bounded by arbitrary non-material interfaces, are developed. Consider a river flow
represented by the movement of a continuum mixture of a fluid and solid with
density q(x, y, z, t), where (x, y, z) are the Cartesian coordinates and t is the time.
The flow is therefore bounded by two general interface surfaces representing the
free surface (subscript s) and the terrain surface (subscript b for bed), given by the
mathematical statements z = zs(x, y, t) and z = zb(x, y, t), respectively (Fig. 1.5).
The velocity components in the (x, y, z) directions are (u, v, w).
With the mixture velocity V = (u, w, v) as the barycentric velocity (Iverson
2005; Wu 2008), the mass conservation equation for the mixture is
Fig. 1.5 3D open channel flow in a river a longitudinal section, b transverse section, c plan view
@ @ @ 2 @ @syx @syy @syz
ðqvÞ þ ðquvÞ þ qv þ ðqvwÞ ¼ þ þ ; ð1:3Þ
@t @x @y @z @x @y @z
@ @ @ @ 2 @szx @szy @szz
ðqwÞ þ ðquwÞ þ ðqvwÞ þ qw ¼ qg þ þ :
@t @x @y @z @x @y @z
ð1:4Þ
Here g is the gravity acceleration and sij the stress tensor, with (i, j) = (x, y, z),
which is symmetric, i.e. sij = sji. Subscript i indicates the axis along which the
stress acts, and j is the axis normal to the plane containing the stress. Moreover, sij
is here introduced with the notation used in the environmental context, i.e., it is the
negative value of the common stress tensor definition used in civil engineering
(Iverson 2005). Equations (1.1)–(1.4) provide the starting point to generate a
depth-integrated model.
10 1 Fundamental Equations of Free Surface Flows
DF @F
¼ þ U rF ¼ 0; ð1:6Þ
Dt @t
where U is the velocity vector representing the deformation of the river bed. Note
that the velocity of displacement of an interface is not necessarily equal to that of
the particles that are momentarily lying upon it. In general, the velocity of the
displacement of a surface is given by the kinematic statement
U ¼ Vb Mb n: ð1:7Þ
Here Vb = (ub, vb, wb) is the fluid mixture velocity on the surface, Mb is the net
volume of fluid mixture crossing normal the interface “b” per unit area and time,
and n is the unit vector normal to the interface. Inserting Eq. (1.7) into Eq. (1.6)
produces
DF @F @F
¼ þ ðVb Mb nÞ rF ¼ þ Vb rF Mb jrF j
Dt @t @t
ð1:8Þ
@F @F @F @F
¼ þ ub þ vb þ wb Mb jrF j ¼ 0:
@t @x @y @z
or
where
" 2 #1=2
@zb 2 @zb
nb ¼ þ þ1 : ð1:14Þ
@x @y
Equation (1.13) is the general kinematic boundary condition at the river bed. This
equation must be preserved in depth-integrated models, regardless of using a
non-slip or slip (ub = vb = wb = 0) assumption at the river bed. A similar equation
applies for the free surface, resulting in
where
" 2 #1=2
@zs 2 @zs
ns ¼ þ þ1 : ð1:16Þ
@x @y
12 1 Fundamental Equations of Free Surface Flows
Zzs
@q @ ðquÞ @ ðqvÞ @ ðqwÞ
þ þ þ dz ¼ 0: ð1:17Þ
@t @x @y @z
zb
Zzs Zzs
@q @ @zs @zb
dz ¼ qdz qs þ qb ; ð1:18Þ
@t @t @t @t
zb zb
Zzs Zzs
@ ðquÞ @ @zs @zb
dz ¼ qudz qs us þ qb ub ; ð1:19Þ
@x @x @x @x
zb zb
Zzs Zzs
@ ðqvÞ @ @zs @zb
dz ¼ qvdz qs vs þ qb vb ; ð1:20Þ
@y @y @y @y
zb zb
Zzs
@ ðqwÞ
dz ¼ qs ws qb wb : ð1:21Þ
@z
zb
or,
Inserting the general kinematic boundary conditions Eqs. (1.13) and (1.15) into
Eq. (1.23) yields
Zzs
@ @ 2 @ @
ðquÞ þ qu þ ðquvÞ þ ðquwÞ dz
@t @x @y @z
zb
Zzs
@sxx @sxy @sxz
¼ þ þ dz: ð1:25Þ
@x @y @z
zb
Zzs Zzs
@ @ @zs @zb
ðquÞdz ¼ qudz qs us þ qb u b ; ð1:26Þ
@t @t @t @t
zb zb
Zzs Zzs
@ 2 @ @zs @zb
qu dz ¼ qu2 dz qs u2s þ qb u2b ; ð1:27Þ
@x @x @x @x
zb zb
14 1 Fundamental Equations of Free Surface Flows
Zzs Zzs
@ @ @zs @zb
ðquvÞdz ¼ quvdz qs us vs þ qb u b v b ; ð1:28Þ
@y @y @y @y
zb zb
Zzs
@
ðquwÞdz ¼ qs us ws qb ub wb ; ð1:29Þ
@z
zb
Zzs Zzs
@sxx @ @zs @zb
dz ¼ sxx dz ðsxx Þs þ ðsxx Þb ; ð1:30Þ
@x @x @x @x
zb zb
Zzs Zzs
@sxy @ @zs @zb
dz ¼ sxy dz sxy s þ sxy b ; ð1:31Þ
@y @y @y @y
zb zb
Zzs
@sxz
dz ¼ ðsxz Þs ðsxz Þb : ð1:32Þ
@z
zb
Equations (1.26)–(1.32) form a set of identities that are substituted into Eq. (1.25),
transforming it into
ð1:33Þ
1.2 General Depth-Integrated Equations over a 3D Terrain 15
ð1:35Þ
ð1:36Þ
Zzs
@ @ @ @ 2
ðqwÞ þ ðquwÞ þ ðqvwÞ þ qw dz
@t @x @y @z
zb
Zzs Zzs
@szx @szy @szz
¼ g qdz þ þ dz ð1:37Þ
@x @y @z
zb zb
16 1 Fundamental Equations of Free Surface Flows
Zzs Zzs
@ @ @zs @zb
ðqwÞdz ¼ qwdz qs ws þ qb wb ; ð1:38Þ
@t @t @t @t
zb zb
Zzs Zzs
@ @ @zs @zb
ðquwÞdz ¼ quwdz qs us ws þ q b ub w b ; ð1:39Þ
@x @x @x @x
zb zb
Zzs Zzs
@ @ @zs @zb
ðqvwÞdz ¼ qvwdz qs vs ws þ q b vb w b ; ð1:40Þ
@y @y @y @y
zb zb
Zzs
@ 2
qw dz ¼ qs w2s qb w2b ; ð1:41Þ
@z
zb
Zzs Zzs
@szx @ @zs @zb
dz ¼ szx dz ðszx Þs þ ðszx Þb ; ð1:42Þ
@x @x @x @x
zb zb
Zzs Zzs
@szy @ @zs @zb
dz ¼ szy dz szy s þ szy b ; ð1:43Þ
@y @y @y @y
zb zb
Zzs
@szz
dz ¼ ðszz Þs ðszz Þb ; ð1:44Þ
@z
zb
Equations (1.24), (1.35), (1.36), and (1.46) are the general non-hydrostatic
depth-integrated mass and momentum conservation equations for mixture flows.
This set of generalized equations accounts for density variations, movable beds,
flows across the free surface and non-hydrostatic stresses over a 3D terrain. The
flow can be turbulent or laminar, steady or unsteady, rotational or irrotational,
uniform or non-uniform, and gradually varied or rapidly varied. The bed may be
rigid or erodible, pervious or impervious, with steep or small slopes, and the free
surface a material surface or receive flows. The fluid may be compressible or
incompressible, with sediments in suspension or clear water (Yen 1973, 1975).
With the general stress tensor considered, the model also applies to avalanche
dynamics using the Mohr–Coulomb yield criterion (Iverson 2005). Open channel
flows are turbulent; Reynolds decomposition is employed in these flows, splitting
the flow variables into mean (denoted by bars) and fluctuating (denoted by primes)
components, e.g., u ¼ u þ u0 ; v ¼ v þ v0 ; w ¼ w þ w0 ; p ¼ p þ p0 ; q ¼ q þ q0 . By
definition, the time-averaged value of a fluctuating component is zero. Employing
these transformations, doing a time-averaging of the 3D equations, and neglecting
viscous contributions, the symmetric stress tensor is (Rodi 1980; White 1991, 2009)
Here r denotes the turbulent Reynolds stress due to time averaging of the Navier–
Stokes equations for fluid flow. The velocity field corresponding to this stress tensor
is a time-averaged mathematical approach representing the mean turbulent motion.
The Reynolds stresses are in reality terms originating from the acceleration, and,
thus, are “apparent stresses.”
With this set of generalized equations, including variable density, turbulence,
and non-hydrostatic pressures, simplified versions will be presented in the ensuing
developments.
18 1 Fundamental Equations of Free Surface Flows
Ben Chie Yen was born on April 14, 1935, at Guangzhou, China, and passed
away aged 66 years on December 23, 2001, at St. Louis IL, USA. He
obtained his BS degree from the National Taiwan University in 1956, the
civil engineer degree from University of Iowa in 1959, and earned his Ph.D.
title there in 1965. After a short stay at Princeton University, he was from
1966 to 1976 Assistant and Associate Professor at the University of Illinois,
Urbana, and from 1976 there Professor of civil engineering. Yen visited later
a number of universities, including University of Karlsruhe in 1974, EPFL
Lausanne in 1982, and National Taiwan University in 1983. He passed away
following cardiovascular aneurysm.
Yen’s Ph.D. thesis was in open channel flow, a topic that he pursued until
the 1970s. He was able to distinguish between the energy and the momentum
equations, resulting in two different formulations with the appropriate aver-
aging coefficients and corresponding loss slopes. He then turned to the
stormwater technology thereby investigating a number of its elements. He, for
instance, pointed to the significant energy losses induced by junction man-
holes, a fact that was often neglected. Yen chaired the second Conference on
Urban Drainage in 1982. He recognized the need to merge water quality and
quantity experts by forming a new international group, the Joint Committee
on Urban Storm Drainage JCUSD, which he chaired also from 1982. He
further investigated unsteady sewer flows and eventually developed into an
expert in this field. He was awarded the Hunter Rouse Hydraulic Engineering
Lecture from the American Society of Civil Engineers ASCE in 1999 for “his
fundamental work on open channel and flow resistance.” He was further
promoted during the 29th IAHR Congress held in Beijing to Honorary
Member of IAHR.
1.3 Saint-Venant Theory 19
In this section, the shallow water equations (SWE) for 2D flows of clear water
(q = const.) over a horizontal plane are derived. If the flow depth h(x, y, t) is smaller
than the characteristic length in the (x, y)-plane, a scaling analysis reveals that,
except for the near-bed boundary layer, the velocity components u and v can be
assumed constant across h and equal to their depth-averaged values U and
V (Liggett 1994). Therefore,
Zzs
1
uðx; y; z; tÞ U ðx; y; tÞ ¼ udz; ð1:48Þ
h
zb
Zzs
1
vðx; y; z; tÞ V ðx; y; tÞ ¼ vdz: ð1:49Þ
h
zb
@h @ @
þ ðUhÞ þ ðVhÞ ¼ 0: ð1:50Þ
@t @x @y
Using the approximations stated above in Eq. (1.35), and neglecting stresses at the
free surface, yields the depth-averaged x-momentum as
@ @ 2 @
ðUhÞ þ U h þ ðUVhÞ
@t @x @y
2 3
Zzs Zzs ð1:51Þ
14@ @ @zb @zb
¼ sxx dz þ sxy dz þ ðsxx Þb þ sxy b ðsxz Þb 5:
q @x @y @x @y
zb zb
This needs further considerations about the stress tensor. Consider first the impli-
cations of using Eqs. (1.48)–(1.49) on the x-momentum balance. Basically, it
amounts to assume that the Boussinesq velocity correction coefficients (Liggett
1994; Katopodes 2019), e.g.,
20 1 Fundamental Equations of Free Surface Flows
Rh Rh Rh
u2 dg v2 dg uvdg
bx ¼ 0
; by ¼ 0
; bxy ¼ 0
; ð1:52Þ
U2h V 2h UVh
are equal to unity, with η as the vertical distance above the channel bed. In uniform
and one-dimensional gradually varied turbulent open channel flows, the velocity
distribution is typically approximated by the logarithmic law of the wall (Fig. 1.6a),
which for a rough bed is (White 1991, Montes 1998)
u 1 g
¼ ln þ B: ð1:53Þ
u j ks
Here u* = (sb/q)1/2 is the shear velocity, j (=0.41) the von Kármán constant, ks the
equivalent roughness height and B (8.5) a constant of integration. Other velocity
distributions as the power law and the wall-wake law are also used in open channel
flow problems (Montes 1998; Chanson 2004).
The Boussinesq velocity correction coefficient is of the order of bx 1.05 for
these flows (Chow 1959; Henderson 1966; Chaudhry 2008). Thus, given the small
contribution of the differential advection originating from the non-uniformity of
u and v with depth, the Boussinesq velocity correction coefficients in the x- and
y-directions are assumed to be unity. Likewise, Eq. (1.39) then reads for the
depth-averaged y-momentum balance
Fig. 1.6 Examples of velocity distributions in open channel flows a log-law velocity distribution
in rough-bed uniform open channel flow, b velocity distribution in hydraulic jumps, with a jet flow
and a roller flow above
1.3 Saint-Venant Theory 21
@ @ 2 @
ðVhÞ þ V h þ ðVUhÞ
@t @y @x
2 3
Zzs Zzs ð1:54Þ
14@ @ @zb @zb
¼ syy dz þ sxy dz þ syy b þ sxy b syz b 5:
q @y @x @y @x
zb zb
Caution is claimed here by pointing out that taking the velocity correction coeffi-
cients equal to unity is a legitimate approximation for gradually varied flows
(setting aside compound channel flows), but not so in rapidly varied flows.
A prominent example is the hydraulic jump (Fig. 1.4b). Within the region of high
turbulence and shear, the u-velocity profiles are highly non-uniform in the vertical
direction. The correction coefficients are easily of the order of bx = 2.5 (Fig. 1.6b),
encompassing a jet-like flow which spreads below a recirculation vortex or roller
(Khan and Steffler 1996a; Castro-Orgaz and Hager 2009). Obviously, these coef-
ficients cannot be neglected. As will be shown in Chap. 5, the SWE permit to catch
hydraulic jumps as local discontinuities, excluding the detailed variation of the field
variables across it, including the free surface and the velocity profiles. This fact has
double implications: (1) the SWE can detect a hydraulic jump, and identify it as a
point-like discontinuity, but cannot resolve the flow within it; (2) there is no need to
introduce any correction coefficient b given that the equations do not resolve the
internal flow phenomena of a hydraulic jump. This feature must be fully accounted
for only if a higher-order depth-averaged model is considered for resolving the
detailed free surface profile of the hydraulic jump (Khan and Steffler 1996a;
Castro-Orgaz and Hager 2009).
Using Eqs. (1.47) for the stress tensor sij, and neglecting stresses of
depth-averaging in Eqs. (1.51) and (1.54), produces
@ @ 2 @
ðUhÞ þ U h þ ðUVhÞ
@t @x @y
2 3
Zzs ð1:55Þ
14@ @zb @zb
¼ pdz þ pb rxy b þ ðrxz Þb 5;
q @x @x @y
zb
@ @ 2 @
ðVhÞ þ V h þ ðVUhÞ
@t @y @x
2 3
Zzs ð1:56Þ
14@ @zb @zb 5
¼ pdz þ pb rxy b þ ryz b :
q @y @y @x
zb
A typical assumption in the derivation of the gradually varied SWE is that the
pressure distribution is hydrostatic. From Eq. (1.4), the vertical momentum balance
is thus for a hydrostatic flow
22 1 Fundamental Equations of Free Surface Flows
dp
¼ qg: ð1:57Þ
dz
Integrating in the z-direction, taking zero pressure as reference at the free surface,
yields
g g
p ¼ qgh 1 ¼ pb 1 : ð1:58Þ
h h
This is a linear distribution of fluid pressure (Fig. 1.7), with the bed pressure equal
to the weight of the water column above, that is,
pb ¼ qgh: ð1:59Þ
To highlight the conditions under which this approximation is valid, consider the
general Eq. (1.46) for the z-momentum balance. This is an equation for the bottom
pressure, which yields, by using Eq. (1.47) for the stress tensor, after assuming that
(u, v, w) are approximated by the corresponding depth-averaged values (U, V, W),
material surfaces at the free surface and the bed (Ms = Mb = 0), depth-averaged
density, zero free surface stresses, and neglecting turbulent stresses due to
depth-averaging,
@ @ @ @zb
pb ¼ qgh þ ðqWhÞ þ ðqWUhÞ þ ðqWVhÞ þ ðrzz Þb ðrxz Þb
@t @x @y @x
@zb
ryz b : ð1:60Þ
@y
@ @ 2 @
ðUhÞ þ U h þ ðUVhÞ
@t @x @y
ð1:61Þ
@h @zb 1 @zb 1
¼ gh gh þ rxy b ðrxz Þb ;
@x @x q @y q
@ @ 2 @
ðVhÞ þ V h þ ðVUhÞ
@t @y @x
ð1:62Þ
@h @zb 1 @zb 1
¼ gh gh þ rxy b ryz b :
@y @y q @x q
To be coherent with Eq. (1.59), the bed slopes in the x- and y-directions are
assumed to be small, eliminating the corresponding stress contributions due to
slopes, resulting in the depth-averaged SWE or Saint-Venant equations
(Vreugdenhil 1994; Toro 2001)
@ @ 2 @ @zs 1
ðUhÞ þ U h þ ðUVhÞ ¼ gh ðrxz Þb ; ð1:63Þ
@t @x @y @x q
@ @ 2 @ @zs 1
ðVhÞ þ V h þ ðVUhÞ ¼ gh ryz b ; ð1:64Þ
@t @y @x @y q
where the gradients of zs = zb + h, the free surface elevation, are used. Finally, with
Cf as a bed friction coefficient, closure for the bed turbulent stresses is given by
1=2 1=2
ðrxz Þb ¼ qCf U U 2 þ V 2 ; ryz b ¼ qCf V U 2 þ V 2 : ð1:65Þ
24 1 Fundamental Equations of Free Surface Flows
Zyr
@h @ @
þ ðUhÞ þ ðVhÞ dy ¼ 0: ð1:66Þ
@t @x @y
yl
Note that within the flow section both h and zb are variable, but zs is a constant.
Now, the following identity is developed by application of Leibniz’s rule:
Zyr Zyr
@h @ @yr @yl @A @yr @yl
dy ¼ hðx; y; tÞdy hr þ hl ¼ hr þ hl ; ð1:67Þ
@t @t @t @t @t @t @t
yl yl
Zyr Zyr
@ ðUhÞ @ @yr @yl @Q @yr @yl
dy ¼ Uhdy ðUhÞr þ ðUhÞl ¼ ðUhÞr þ ðUhÞl ;
@x @x @x @x @x @x @x
yl yl
ð1:69Þ
Zyr
@ ðVhÞ
dy ¼ ðVhÞr ðVhÞl : ð1:71Þ
@y
yl
To further simplify Eq. (1.72), Chen and Chow (1968) assumed that the
depth-averaged velocity field (U, V) must satisfy a kinematic boundary condition at
the banks, whereas Jain (2001) considered the usual case of a cross section of zero
water depth at the banks, e.g., zs = zb (Fig. 1.9). One may state, therefore,
26 1 Fundamental Equations of Free Surface Flows
@y
@yl @yl þ Ul @y
@x Vl ¼ 0 ) Option 1
l l
hl þ Ul Vl ¼ 0 ) @t ð1:73Þ
@t @x hl ¼ 0 ) Option 2
@y
@yr @yr þ Ur @y
@x Vr ¼ 0 ) Option 1
r r
hr þ Ur Vr ¼ 0 ) @t ð1:74Þ
@t @x hr ¼ 0 ) Option 2
By using either of the two arguments, Eq. (1.72) simplifies to the section-averaged
continuity equation
@A @Q
þ ¼ 0: ð1:75Þ
@t @x
Further, the x-momentum equation [Eq. (1.51)] will be laterally integrated. For
convenience in ensuing developments, the identity
Zzs Zzs
@sxy @sxz @ @zb
þ dz ¼ sxy dz þ sxy b ðsxz Þb ; ð1:76Þ
@y @z @y @y
zb zb
ð1:77Þ
ð1:78Þ
Assuming hydrostatic pressure distribution and neglecting rxx, one may write
Eq. (1.78) as
2 3
Zzs
@ @ 2 @ 14@ 1 @zb @rxy @rxz
ðUhÞ þ U h þ ðUVhÞ ¼ qgh þ qgh
2
þ dz5;
@t @x @y q @x 2 @x @y @z
zb
ð1:79Þ
1.3 Saint-Venant Theory 27
or,
2 z 3
Zs
@ @ 2 @ @zs 14 @rxy @rxz
ðUhÞ þ U h þ ðUVhÞ ¼ gh þ þ dz5:
@t @x @y @x q @y @z
zb
ð1:80Þ
ð1:81Þ
The following identities are generated to integrate the acceleration terms [left-hand
side of Eq. (1.81)]
Zyr Zyr
@ @ @yr @yl
ðUhÞdy ¼ Uhdy ðUhÞr þ ðUhÞl ; ð1:82Þ
@t @t @t @t
yl yl
Zyr Zyr
@ ð U 2 hÞ @ @yr 2 @yl
dy ¼ U 2 hdy U 2 h r þ U h l ; ð1:83Þ
@x @x @x @x
yl yl
Zyr
@ ðUVhÞ
dy ¼ ðUVhÞr ðUVhÞl : ð1:84Þ
@y
yl
Zyr
@ @ 2 @ @Q @ Q2
ðUhÞ þ U h þ ðUVhÞ dy ¼ þ
@t @x @y @t @x A
yl ð1:85Þ
@yl @yl @yr @yr
þ ðUhÞl þ Ul Vl ðUhÞr þ Ur Vr ;
@t @x @t @x
28 1 Fundamental Equations of Free Surface Flows
Zyr
@ @ 2 @ @Q @ Q2
ðUhÞ þ U h þ ðUVhÞ dy ¼ þ : ð1:86Þ
@t @x @y @t @x A
yl
In this process, the variation of U in the lateral direction was neglected. If the
variation of the velocity component u within the cross section is accounted for, the
momentum flux term Q2/A should be multiplied by the Boussinesq correction
coefficient for the section (Liggett 1994; Montes 1998), given by
R
u2 dA
b¼ A
: ð1:87Þ
ðQ2 =AÞ
For a horizontal water surface within the cross section, the free surface slope term
reads
Zyr Zyr
@zs @zs @zs
g hðx; y; tÞdy ¼ g hðx; y; tÞdy ¼ g A: ð1:88Þ
@x @x @x
yl yl
Using Green’s theorem, the following identity is generated (Keulegan and Patterson
1943)
Zyr Zzs Z I
@rxy @rxz @rxy @rxz
þ dzdy ¼ þ dA ¼ rxy dz rxz dy ;
@y @z @y @z
yl zb A C
ð1:89Þ
where the line integral extends along a contour C as the sum of the wetted perimeter
P (Fig. 1.10) and the free surface width B. The integrand of the line integral is the x-
component of the fluid force exerted on the channel solid contour and free surface.
This shear force is denoted by sb, from which a mean shear stress so is defined by
I I
rxy dz rxz dy ¼ sb dP ¼ so P; ð1:90Þ
C C
This equation is valid for non-prismatic channels and is a form widely used in river
hydraulics computations (Cunge et al. 1980; Wu 2008). Defining the friction slope
Sf as
so P so
Sf ¼ ¼ ; ð1:92Þ
qg A qgRh
where Rh = A/P is the hydraulic radius, the following “slope form” results (Cunge
et al. 1980; Chow et al. 1988)
@Q @ Q2 @h
þ ¼ gA So Sf ; ð1:93Þ
@t @x A @x
where So ¼ @zb =@x is the bottom slope. For practical computations, consider a
flow depth and bed elevation taking as reference the thalweg; thus h = h(x, t) and
zb = zb(x, t) (Fig. 1.11). Consider further a transformation of Eq. (1.93), which is
widely used in practice. The hydrostatic pressure force on an arbitrary cross section
is
Z Z ZhðxÞ
Fp ¼ pdA ¼ qg ðh gÞdA ¼ qg ðh gÞbðx; gÞdg; ð1:94Þ
A A 0
Z ZhðxÞ
Fp
¼ Ah ¼ ðh gÞdA ¼ ðh gÞbðx; gÞdg: ð1:95Þ
qg
A 0
ð1:96Þ
This relation expresses the spatial variation of hydrostatic pressure forces per unit
weight in a non-prismatic channel.
Inserting Eq. (1.96) into Eq. (1.93) produces the alternative form for
non-prismatic channels
ZhðxÞ
@Q @ Q2 @b
þ þ gAh ¼ gA So Sf þ g ðh gÞ dg: ð1:97Þ
@t @x A @x
0
2 ZhðxÞ
@Q @ Q @b
þ b þ gAh ¼ gA So Sf þ g ðh gÞ dg: ð1:98Þ
@t @x A @x
0
If the pressure is assumed to be hydrostatic and the velocity uniform, the specific
momentum M in free surface flows is defined as (Jaeger 1956; Montes 1998)
1.3 Saint-Venant Theory 31
Fig. 1.12 Lateral distribution of depth-averaged velocity component in x-direction for uniform
flow in compound channel cross section, assuming that the velocity is constant in main channel
and flood plains, each linked to a different value of the equivalent roughness ks
Z 2
u p Q2 Q2
M¼ þ dA b þ Ah þ Ah: ð1:99Þ
g qg gA gA
A
The definition of M arises thus naturally from the application of the x-momentum
balance, as evidenced on inspecting Eq. (1.97). For a prismatic channel, Eq. (1.98)
reduces to
@Q @ Q2
þ þ gAh ¼ gA So Sf : ð1:100Þ
@t @x A
or,
Q2 dA dh dzb
¼ gA gA gASf : ð1:103Þ
A2 dx dx dx
For a non-prismatic channel, the x-derivative of the flow area A can be expressed as
dA @A dh @A
¼ þ : ð1:104Þ
dx @h dx @x h¼const:
These partial derivatives are trivially simple to evaluate in man-made channels; for
example, the result for a rectangular channel of variable width is
@A db @A
Aðx; hÞ ¼ bð xÞh ) ¼h ; ¼ b: ð1:105Þ
@x h¼const: dx @h
from which the ordinary differential equation describing the flow surface profile is
(Le Méhauté 1976; Katopodes 2019)
Q2 @A 2
dh So Sf þ gA3 @x h¼const: So Sf þ FB @A @x h¼const:
¼ ¼ : ð1:107Þ
dx Q2
1 gA 3 B 1 F2
Q=A
F¼ ; ð1:108Þ
ðgA=BÞ1=2
1.3 Saint-Venant Theory 33
Zh
@A @
¼ bðx; gÞdg ¼ B: ð1:109Þ
@h @h
0
dh So Sf
¼ : ð1:110Þ
dx 1F2
Equations (1.107) and (1.110) are widely used in open channel hydraulics, but it
should be borne in mind that they are unreliable in highly curvilinear flows, as for
flows over a dam spillway crest operating at a high head (Montes 1998).
Steffler and Jin (1993) and Khan and Steffler (1996b, c) developed a depth-averaged
non-hydrostatic 1D momentum model for flows in the vertical plane (x, z), to be
described here. Other approximations are presented in Chap. 11. Consider a 2D
steady flow over a wavy bed (Fig. 1.14a). The velocity component u in the x-
direction is approximated by its depth-averaged value U, like in the Saint-Venant
theory, but the vertical velocity w and the non-hydrostatic fluid pressure p are
assumed to vary linearly in the vertical direction (Fig. 1.14b). Assuming 1D flow in
the x-direction, using Eq. (1.47) for the stress tensor and neglecting depth-averaged
turbulent stresses, Eq. (1.51) takes the form
0 1
Zzs
@ @ @ 2 p A 1 @zb @zb
ðUhÞ þ U hþ dz ¼ pb ðrxz Þb þ rxx ; ð1:111Þ
@t @x q q @x @x
zb
@ @ @zb
pb ¼ qgh þ ðqWhÞ þ ðqWUhÞ þ ðrzz Þb ðrxz Þb : ð1:112Þ
@t @x @x
34 1 Fundamental Equations of Free Surface Flows
Fig. 1.14 1D
non-hydrostatic flow over
uneven terrain a longitudinal
profile, b assumed horizontal
and vertical velocity, and
pressure profiles
Simple shear along the bed slope with tangential stress sb is considered. For this
choice, the bed Reynolds stresses are given by (Steffler and Jin 1993; Castro-Orgaz
and Hager 2017)
where h is the bed slope angle. It is therefore obvious that the bed Reynolds stresses
are not negligible on steep slopes. For h ! 0 the channel is of a small slope,
resulting in
This is an approximation used in the Saint-Venant theory, but not here. The
Reynolds stresses contribution in the x-momentum equation is thus
@zb sin h
ðrxx Þb þ ðrxz Þb ¼ 2sb cos h sin h þ sb cos2 h sin2 h
@x cos h ð1:115Þ
¼ sb cos2 h þ sin2 h ¼ sb :
1.4 Non-hydrostatic Theory 35
Inserting Eqs. (1.115)–(1.116) for a pure bed shear into Eqs. (1.111) and (1.112)
gives
0 1
Zzs
@ @ @ 2 p A pb @zb sb
ðUhÞ þ U hþ dz ¼ ; ð1:117Þ
@t @x q q @x q
zb
@ @ @zb
pb ¼ qgh þ q ðWhÞ þ q ðWUhÞ þ sb : ð1:118Þ
|{z} @t @x @xffl}
hydrostatic term
|fflfflfflfflffl{zfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} |fflffl{zffl
local acceleration convective acceleration Reynolds stresses
The vertical velocity w is assumed to vary linearly with depth (Fig. 1.14b). Its
depth-averaged value W is thus given by
ws þ wb
W¼ ; ð1:119Þ
2
where the free surface and bed kinematic boundary conditions are, respectively,
@ @
ws ¼ ðh þ zb Þ þ U ðh þ zb Þ; ð1:120Þ
@t @x
@zb @zb
wb ¼ þU : ð1:121Þ
@t @x
A distribution p = p(z) must be introduced into Eq. (1.117) for model closure.
A linear distribution is assumed following Khan and Steffler (1996b) and Denlinger
and Iverson (2004)
g g
p ¼ pb 1 ¼ ðp1 þ qghÞ 1 ; ð1:122Þ
h h
where p1 is the difference between the actual pb and the hydrostatic bottom pressure
(qgh). The non-hydrostatic pressure force is thus
36 1 Fundamental Equations of Free Surface Flows
Zzs Zh
h2 hp1
pdz ¼ pdg ¼ qg þ : ð1:123Þ
2 2
zb 0
or
@ @ h pb pb @zb sb
ðUhÞ þ U hþ
2
¼ : ð1:125Þ
@t @x 2q q @x q
dM pb dzb sb
¼ ; ð1:126Þ
dx q dx q
h pb
M ¼ U2h þ ; ð1:127Þ
2q
d dzb
pb ¼ qgh þ q ðWUhÞ þ sb : ð1:128Þ
dx dx
ws þ wb
W¼ : ð1:129Þ
2
d
ws ¼ U ðh þ zb Þ; ð1:130Þ
dx
dzb
wb ¼ U : ð1:131Þ
dx
Consider uniform flow on a steep slope (Fig. 1.15), for which the flow depth is a
constant and there are not spatial variations of any flow variable. Equations (1.126)
and (1.128) simplify then to
Combining Eqs. (1.134) and (1.135) results in the non-hydrostatic bottom pressure
on a steep slope as (Chaudhry 2008; Castro-Orgaz et al. 2015; Castro-Orgaz and
Hager 2017)
Fig. 1.15 1D
non-hydrostatic uniform flow
on a steep slope
38 1 Fundamental Equations of Free Surface Flows
pb h
¼ ¼ hcos2 h: ð1:136Þ
qg 1 þ tan2 h
pb h p1
¼ ¼ þ h; ð1:137Þ
qg 1 þ tan2 h qg
Fig. 1.16 1D
non-hydrostatic gradually
varied flow on a steep slope
a vertical pressure profile,
b free surface and bottom
pressure head line
1.4 Non-hydrostatic Theory 39
The vertical pressure distribution is therefore linear with this value of p1 for the
non-hydrostatic correction (Fig. 1.16a). The specific momentum M is thus
2
dzb
2 2
h hp1 h dx h2 0h
2
M ¼ U2h þ g þ ¼ U2h þ g g 2 ¼ U h þ g ; ð1:139Þ
2
2 2q 2 dzb 2 2
1þ
dx
where g0 is the enhanced gravity to account for slope effects (Denlinger and Iverson
2004; Denlinger and O’Connell 2008), given by
1
g0 ¼ g 2 : ð1:140Þ
dzb
1þ
dx
Inserting Eqs. (1.139) and (1.141) into Eq. (1.126) produces the gradually varied
flow equation on a steep slope as
d h2 dzb sb
U 2 h þ g0 ¼ g0 h : ð1:142Þ
dx 2 dx q
With U = q/h as the depth-averaged horizontal velocity, Eq. (1.142) transforms into
dh q2 dzb sb
1 0 3 ¼ : ð1:143Þ
dx gh dx qg0 h
The stress sb acts tangentially to the sloping plane (Steffler and Jin 1993)
(Fig. 1.16b). Therefore, it is parametrized as (Khan and Steffler 1996b)
with Cf as the bed friction coefficient and Uo as the velocity component parallel to
the sloping plane. The velocity components of the Boussinesq-type model [given
by Eqs. (1.126–1.131)] in the (x, z) directions are (U, W). The following approx-
imation is now used (Khan and Steffler (1996b)
40 1 Fundamental Equations of Free Surface Flows
1=2
Uo U 2 þ W 2 : ð1:145Þ
Using this relation for gradually varied flows, the coupling of Eqs. (1.145)–(1.146)
generates
" 2 #1=2
dzb
Uo ¼ U 1 þ : ð1:147Þ
dx
Thus, combining Eqs. (1.144) and (1.147) yields the bottom shear stress as
" 2 #
dzb
sb ¼ qCf U 1 þ
2
; ð1:148Þ
dx
1=2
Equation (1.143) can be rewritten with F ¼ q=ðgh3 Þ as
" 2 #
dzb dzb
Sf 1 þ
dh dx dx
¼ " # : ð1:150Þ
dx dzb 2
1 F 1þ
2
dx
This is the gradually varied flow equation for non-hydrostatic flow on steep rect-
angular channels. It will be considered in Chap. 3.
1.5 Sediment Transport and Movable Beds 41
The free surface is taken as a material surface, and the bed is considered erodible,
but a non-slip kinematic boundary conditions is implemented, that is, u = v =
w = 0 at the bed (Wu 2008). Equation (1.152) reduces then to
Taking average values of u, v, namely (U, V), and q, denoted by the same symbol to
simplify notation, Eq. (1.153) simplifies to
@ @ @ @zb
ðqhÞ þ ðqUhÞ þ ðqVhÞ þ qb ¼ 0: ð1:154Þ
@t @x @y @t
Note that the movable bed due to sediment transport is accounted for by inclusion
of the time derivative @zb =@t in the continuity equation. An average value of
density is used, but a Reynolds decomposition of this variable into mean and
Fig. 1.17 One-dimensional flow with movable bed and sediment transport
42 1 Fundamental Equations of Free Surface Flows
fluctuating parts is not done here for the sake of simplicity. Lateral integration of
Eq. (1.154) from the left to the right channel bank yields
Zyr
@ @ @ @zb
ðqhÞ þ ðqUhÞ þ ðqVhÞ þ qb dy ¼ 0: ð1:155Þ
@t @x @y @t
yl
Zyr Zyr
@ @ @yr @yl
ðqhÞdy ¼ qhdy ðqhÞr þ ðqhÞl
@t @t @t @t
yl yl
@ @yr @yl
¼ ðqAÞ ðqhÞr þ ðqhÞl ; ð1:156Þ
@t @t @t
Zyr Zyr
@ ðqUhÞ @ @yr @yl
dy ¼ qUhdy ðqUhÞr þ ðqUhÞl
@x @x @x @x
yl yl ð1:157Þ
@ @yr @yl
¼ ðqQÞ ðqUhÞr þ ðqUhÞl ;
@x @x @x
Zyr
@ ðqVhÞ
dy ¼ ðqVhÞr ðqVhÞl ; ð1:158Þ
@y
yl
Zyr Zyr
@zb @ @yr @yl
qb dy ¼ qb þ qb ðzb Þl
zb dy qb ðzb Þr
@t @t @t @t
yl yl ð1:159Þ
@Ab @yr @yl @Ab @B
¼ qb qb z s ¼ qb qb zs ;
@t @t @t @t @t
where Ab is the bed area above the reference datum (Fig. 1.17). Summing
Eqs. (1.156)–(1.159) generates
@ @ @yl @yl @yr @yr
ðqAÞ þ ðqQÞ þ hl þ Ul Vl qhr þ Ur Vr
@t @x @t @x @t @x
@Ab @B
þ qb q b zs ¼ 0:
@t @t
ð1:160Þ
Using Eqs. (1.73)–(1.74), and dropping the term related to @B=@t; which is not
considered in river flow models for fast geomorphic flows (Wu 2008), Eq. (1.160)
yields finally
1.5 Sediment Transport and Movable Beds 43
@ @ @Ab
ðqAÞ þ ðqQÞ þ qb ¼ 0: ð1:161Þ
@t @x @t
This simplifies upon assuming that the free surface is a material surface and using a
non-slip velocity at the bed, to
@ @ 2 @
ðqUhÞ þ qU h þ ðqUVhÞ
@t @x @y
2 3
Zzs Zzs ð1:165Þ
@ @ @z @zb
¼ 4 ðsxz Þb 5:
b
sxx dz þ sxy dz þ ðsxx Þb þ sxy b
@x @y @x @y
zb zb
Zzs Zzs
@sxy @sxz @ @zb
þ dz ¼ sxy dz þ sxy b ðsxz Þb ; ð1:166Þ
@y @z @y @y
zb zb
ð1:167Þ
The stress tensor given by Eqs. (1.47) is used, and again noted that even though the
velocity field (u, v, w) represents the time-averaged velocity field when this stress
tensor is used, we do not conduct here a Reynolds decomposition of density for
simplicity’s sake. Once the stress tensor is introduced into Eq. (1.167), neglecting
normal turbulent stresses and assuming a hydrostatic pressure distribution, the
resulting equations is
Zzs
@ @ 2 @ @
ðqUhÞ þ qU h þ ðqUVhÞ ¼ qgðzs zÞdz
@t @x @y @x
zb
Zzs
@zb @rxy @rxz
qgh þ þ dz;
@x @y @z
zb
ð1:168Þ
1.5 Sediment Transport and Movable Beds 45
or,
@ @ 2 @ @ 1
ðqUhÞ þ qU h þ ðqUVhÞ ¼ qgh2
@t @x @y @x 2
Zzs
@zb @rxy @rxz
qgh þ þ dz:
@x @y @z
zb
ð1:169Þ
@ @ 2 @ @h 1 @q
ðqUhÞ þ qU h þ ðqUVhÞ ¼ qgh gh2
@t @x @y @x 2 @x
Zzs
@zb @rxy @rxz
qgh þ þ dz;
@x @y @z
zb
ð1:170Þ
and using the gradient of the free surface, Eq. (1.170) transforms into
@ @ 2 @ @zs
ðqUhÞ þ qU h þ ðqUVhÞ ¼ qgh
@t @x @y @x
Zzs
1 @q @rxy @rxz
gh2 þ þ dz:
2 @x @y @z
zb
ð1:171Þ
This is a suitable form of the depth-integrated x-momentum equation for flows with
sediment transport over movable beds. Lateral integration of Eq. (1.171) from the
left to the right channel bank yields
Zyr
@ @ 2 @
ðqUhÞ þ qU h þ ðqUVhÞ dy
@t @x @y
yl
2 3
Zyr Zzs
@z 1 @q @r @r
¼ 4qgh dz5dy:
s xy xz
gh2 þ þ ð1:172Þ
@x 2 @x @y @z
yl zb
46 1 Fundamental Equations of Free Surface Flows
Zyr Zyr
@ @ @yr @yl
ðqUhÞdy ¼ qUhdy ðqUhÞr þ ðqUhÞl ; ð1:173Þ
@t @t @t @t
yl yl
Zyr Zyr
@ 2 @ @yr 2 @yl
qU h dy ¼ qU 2 hdy qU 2 h r þ qU h l ; ð1:174Þ
@x @x @x @x
yl yl
Zyr
@
ðqUVhÞdy ¼ ðqUVhÞr ðqUVhÞl : ð1:175Þ
@y
yl
Zyr 2
@ @ 2 @ @ @ Q
ðqUhÞ þ qU h þ ðqUVhÞ dy ¼ ðqQÞ þ q : ð1:176Þ
@t @x @y @t @x A
yl
Zyr Zyr
@zs @zs @zs
g hdy ¼ g hdy ¼ g A: ð1:177Þ
@x @x @x
yl yl
Zyr Zzs
@rxy @rxz
þ dzdy ¼ so P: ð1:178Þ
@y @z
yl zb
ð1:179Þ
2
@ @ Q @zs 1 @q
ðqQÞ þ q ¼ qgA so P g hp A; ð1:180Þ
@t @x A @x 2 @x
or
2
@ @ Q @zs 1 @q
ðqQÞ þ q ¼ qgA qgASf g hp A: ð1:181Þ
@t @x A @x 2 @x
This is a widely used momentum model for 1D river flows with sediment transport
(Wu 2008). In this equation, the variable fluid density in space and time due to
sediment transport is fully accounted for.
References
Boussinesq, J. (1877). Essai sur la théorie des eaux courantes [Essay on the theory of water flow].
Mémoires présentés par divers savants à l’Académie des Sciences, Paris 23, pp. 1–680
(in French).
Cantero-Chinchilla, F., Castro-Orgaz, O., Dey, S., & Ayuso, J. L. (2016). Nonhydrostatic dam
break flows II: One-dimensional depth-averaged modelling for movable bed flows. Journal of
Hydraulic Engineering, 142(12), 04016069.
Castro-Orgaz, O., & Hager, W. H. (2009). Classical hydraulic jump: Basic flow features. Journal
of Hydraulic Research, 47(6), 744–754.
Castro-Orgaz, O., & Hager, W.H. (2017). Non-hydrostatic free surface flows. In Advances in
geophysical and environmental mechanics and mathematics (696 pages), Berlin: Springer,
https://doi.org/10.1007/978-3-319-47971-2.
Castro-Orgaz, O., Hutter, K., Giráldez, J. V., & Hager, W. H. (2015). Non-hydrostatic granular
flow over 3D terrain: New Boussinesq-type gravity waves? Journal of Geophysical Research:
Earth Surface, 120(1), 1–28.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). Berlin: Springer.
Chen, C. L., & Chow, V. T. (1968). Hydrodynamics of mathematically simulated surface runoff.
Hydraulic Engineering Series No. 18, Civil Engineering Studies. Urbana-Champaign:
University of Illinois.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Chow, V. T., Maidment, D. R., & Mays, L. W. (1988). Applied hydrology. New York:
McGraw-Hill.
Cunge, J. A., Holly, F. M., & Verwey, A. (1980). Practical aspects of computational river
hydraulics. London: Pitman.
de Saint-Venant, A.B. (1871). Théorie du mouvement non permanent des eaux, avec application
aux crues des rivières et à l’introduction des marrées dans leur lit [Theory of unsteady water
movement, applied to floods in rivers and the effect of tidal flows]. Comptes Rendus de
l’Académie des Sciences, 73, 147–154; 73, 237–240 (in French).
Denlinger, R. P., & Iverson, R. M. (2004). Granular avalanches across irregular three-dimensional
terrain: 1. Theory and computation. Journal of Geophysical Research: Earth Surface, 109(F1),
F01014. https://doi.org/10.1029/2003JF000085.
48 1 Fundamental Equations of Free Surface Flows
2.1 Introduction
@A @Q
þ ¼ 0; ð2:1Þ
@t @x
1 @Q @M
þ ¼ A So Sf : ð2:2Þ
g @t @x
Here, u is the velocity component in the x-direction, p fluid pressure, and h the
depth below the free surface of the centroid of area A. Equation (2.2) is only valid
for prismatic channels. Equations (2.1)–(2.2) are differential equations even for the
simple case of steady gradually varied flow. The conservation laws can be
expressed in integral form if an integration in the flow direction x is accomplished
by considering a control volume of finite length. However, not all hydraulic
problems are solved by the isolated consideration of mass and momentum con-
servations. Energy conservation in integral form needs to be added. Its use is
dQ
¼ 0; ð2:4Þ
dx
dM
¼ A So Sf : ð2:5Þ
dx
The differential form of the energy equation must be considered now. The total
energy head H is defined based on the total energy flow across section A as (Jaeger
1956)
Z
1 u2 þ v 2 þ w 2 p
H¼ þ þ z udA: ð2:6Þ
UA 2g qg
A
Herein, the kinetic energy due to the velocity components (u, v, w), the potential
energy due to elevation z, and the work done by fluid pressure p are accounted for
(Liggett 1994; Montes 1998). Consider gradually varied flow in a channel of small
bottom slope, and assume that the velocity components v = w = 0. Further, assume
that u is uniform across cross section A and equal to the mean value U = Q/A, and
that the vertical pressure distribution follows the hydrostatic law. Under these
simplifications, Eq. (2.6) yields with zb = zb(x) as the bottom elevation relative to
an arbitrary datum (Jaeger 1956; Montes 1998)
Z 2
1 u þ v2 þ w2 p U2 Q2
H¼ þ þ z udA zb þ h þ ¼ zb þ h þ :
UA 2g qg 2g 2gA2
A
ð2:7Þ
This relation is a true energy equation and shall not be mistakenly linked to
Bernoulli’s equation resulting from the integration of the momentum equation
along the streamlines in inviscid and incompressible steady fluid flow (Rouse 1938;
Liggett 1994; Montes 1998). For steady flow with constant discharge Q, the dif-
ferential energy equation obtained from the first law of thermodynamics states that
the variation of H equals the rate of conversion of energy into heat (Liggett 1994;
Montes 1998) (Fig. 2.1)
2.1 Introduction 53
Fig. 2.1 Definition sketch for differential form of energy principle under steady flow
dH
¼ Se ; ð2:8Þ
dx
where Se is the gradient of dissipated energy or the energy head slope. Note that
conceptually Se is different from Sf; the former is a measure of the internal energy
dissipation within a volume of fluid, whereas Sf represents the external shear forces
acting on the boundaries of the control volume.
Equations (2.4), (2.5), and (2.8) are the differential forms of the mass,
momentum, and energy balances for hydrostatic pressure and uniform velocity
distributions (Yen 1973); their integration along a control volume of length (x1–x2)
(Fig. 2.2) yields (Jain 2001)
Q1 Q2 ¼ 0; ð2:9Þ
Zx1
M1 M2 ¼ A So Sf dx; ð2:10Þ
x2
Zx1
H1 H2 ¼ Se dx: ð2:11Þ
x2
Fig. 2.2 Control volume for steady flow in a prismatic channel of arbitrary cross section
54 2 Energy and Momentum Principles
These equations are statements to be satisfied for any flow within a control volume.
Non-hydrostatic pressure and non-uniform velocity distributions could be
accounted for in H and M (Jaeger 1956; Montes 1998; Jain 2001). The boundary
sections are usually positioned in a gradually varied flow zone, allowing for a
simple evaluation of H and M as functions of the local flow depth. Given the
absence of mass sources/sinks (e.g., rain/infiltration), Eq. (2.9) states constant
discharge across the control volume. However, neither M nor H is conserved within
the control volume given the terms related to Sf and Se, acting as sinks of
momentum and energy. Even though this is the general situation, there are many
cases where these terms can be neglected. For example, in a hydraulic jump on a
horizontal bottom, the shear forces are smaller than the pressure forces plus
momentum flux, resulting in M = const. from Eq. (2.10). The energy dissipation is
significant and cannot be overlooked, however. This is an example on how energy
and momentum laws are used as complementary principles: The momentum
equation is used to compute the flow depth h2 at the downstream section of the
jump for known values of Q and h1, whereas Eq. (2.11) is applied to evaluate the
energy dissipation produced by the hydraulic jump.
Q2
H 1 ¼ H 2 ¼ H ¼ zb þ h þ ¼ const: ð2:12Þ
2gA2
The specific energy head E is defined taking the channel bottom as reference. Thus,
Q2
E ¼ hþ : ð2:13Þ
2gA2
It represents therefore the total energy head above the channel bottom at any
section. Inserting Eq. (2.13) into Eq. (2.12) produces
zb ð xÞ þ Eð xÞ ¼ const: ð2:14Þ
This is the equation governing ideal fluid flow across channel transitions, to be used
below. It is a suitable approximation in gradual transitions involving a smooth
2.2 Energy Principle 55
variation of bottom elevation and channel shape, without zones of flow separation
where energy losses have to be accounted for (Rouse 1938; Montes 1998). What
are the properties of Eq. (2.13)? Consider a rectangular channel of width b, with
A = bh, and unit discharge q = Q/b. Equation (2.13) is rewritten as
U2 q2
E ¼ hþ ¼ hþ : ð2:15Þ
2g 2gh2
q2
ðE hÞh2 ¼ ¼ const: ð2:16Þ
2g
This is a cubic in h, plotted in Fig. 2.3. Of the three roots, one is negative, lacking
of physical meaning and therefore discarded. The curve has the asymptotes
(E–h) = 0 and h = 0. The first asymptote is a straight line inclined 45° with respect
to the E-axis. The second asymptote is the E-axis. The three roots of Eq. (2.16) are
determined analytically for a given value of E (Chanson 2004; Jeppson 2011).
Consider a given value of E. If a vertical line is plotted in Fig. 2.3, the inter-
section with the curve E = E(h) yields two points 1 and 2. For a given discharge, a
certain specific energy head is possible for two different values of the flow depth,
namely h1 and h2. These are called alternate depths (Henderson 1966; Jain 2001).
Note from Fig. 2.3 that at point 1, flow depth h corresponds to the dominant part of
E, whereas the velocity head U2/(2g) is small. At point 2, where the value of E is
identical, the trend is reversed: U2/(2g) is the dominant term, whereas h is small.
This observation suggests that each of the alternate depths represents a different
regime, one of small velocity and large flow depths (hydrostatic forces) and the
other representing a flow of small depth at high velocity. The slow velocity regime
is referred to as “subcritical flow,” whereas the high-speed flow regime involves
“supercritical flow.” The limit between these two states is marked by point c in
Fig. 2.3, referred to as the “critical flow condition,” representing the point for which
the value of E is the minimum possible for a given value q. It is impossible to
transport the flow rate q with a value of E below Emin, therefore. The flow depth at
critical condition is called critical (subscript c) flow depth hc. Critical flow is an
important tool in open channel hydraulics and includes important properties, the
first introduced in Fig. 2.3, namely minimum specific energy. From Eq. (2.15), a
quadratic relation between the alternate depths exists.
dE q2
¼ 1 3: ð2:17Þ
dh gh
Setting this expression equal to zero, the extreme of the E-h curve yields the critical
flow condition, namely
q2
1 ¼ 0: ð2:18Þ
gh3c
Inserting Eq. (2.18) into Eq. (2.15), the minimum specific energy head is
q2 1 q2 3
Emin ¼ hc þ ¼ h c 1 þ ¼ hc : ð2:20Þ
2gh2c 2 gh3c 2
The critical depth equals two-thirds of the minimum specific energy, therefore,
2
hc ¼ Emin : ð2:21Þ
3
Uc2 2 U2 Uc2 1 1
Emin ¼ hc þ ¼ Emin þ c ) ¼ Emin ¼ hc : ð2:22Þ
2g 3 2g 2g 3 2
q2 Uc2
¼ ¼ 1; ð2:23Þ
gh3c ghc
or
Uc
F¼ ¼ 1: ð2:24Þ
ðghc Þ1=2
58 2 Energy and Momentum Principles
The Froude number F is unity at the critical flow condition, therefore. From
Fig. 2.3, subcritical flow (e.g., point 1) is characterized by U < Uc and thus F\1.
Likewise, for supercritical flow (e.g., point 2) U > Uc and F [ 1. Using the critical
depth hc as scaling length, Eq. (2.15) is rewritten as
E h 1 h 2
¼ þ : ð2:25Þ
hc hc 2 hc
This is the dimensionless specific energy head curve, plotted in Fig. 2.4. It is a universal
curve for rectangular channel flow, involving the critical flow condition E/hc = 3/2.
There is another interesting feature of Eq. (2.15). Let us rewrite it as
Consider the variation of q with h under constant specific energy head E (Fig. 2.5). Note
that q = 0 either if the flow depth is zero or if E = h. The maximum of the curve follows
from differentiation of Eq. (2.26) with respect to h for E = const., yielding
dq
2q ¼ 4ghðE hÞ 2gh2 ¼ 4ghE 6gh2 : ð2:27Þ
dh
2
hc ¼ E: ð2:28Þ
3
Under the critical flow condition, the discharge is a maximum for the available
value of the specific energy head. This is an important result of great relevance to
establish the head–discharge curve of control structures based on the critical flow
theory (Jaeger 1956; Bos 1976; Ackers et al. 1978; Montes 1998), as shown below.
Another way of expressing Eq. (2.30) is
1=2
qmax ¼ Cd gE 3 ; ð2:31Þ
Consider the position where critical flow is established. Figure 2.6 shows a sketch for
shallow ideal fluid flow over a dam crest. This flow is discussed in Chap. 4, but here
introduced to illustrate an important case of critical flow, following Henderson (1966).
Given that the total energy head H is a constant, Eq. (2.14) is applied. The flow
over the spillway crest changes from upstream subcritical to downstream super-
critical. The detailed computation of the free surface profile is described in Chap. 4.
Differentiation of Eq. (2.14) with respect to x produces
dE dzb dE dh dzb
þ ¼ þ ¼ 0; ð2:33Þ
dx dx dh dx dx
q2 d2 h 3q2 dh 2 d2 zb
1 3 þ þ 2 ¼ 0: ð2:35Þ
gh dx2 gh4 dx dx
1=2
Inserting F ¼ q=ðgh3 Þ gives
d2 h 3 2 dh 2 d2 zb
1F 2
þ F þ 2 ¼ 0: ð2:36Þ
dx2 h dx dx
Let F ¼ 1 for critical flow, the free surface slope at the critical section is (Hager
2010)
1=2
dh hc d2 z b
¼ : ð2:37Þ
dx c 3 dx2 c
Real solutions to Eq. (2.37) result for a bed curvature at the critical section
d2zb/dx2 < 0; e.g., the bed profile must be convex, as is characteristic for a spillway.
For a concave bottom profile imaginary numbers result, excluding a real solution at
a minimum elevation of the bottom profile.
Consider a rectangular horizontal channel with a gradual variation of channel
width. Examples of this structure are flumes used for discharge measurement
(Montes 1998) (Fig. 2.8).
To apply the 1D approach pursued here, the contraction–expansion of the sec-
tions must be gradual, to avoid flow separation and 3D effects. For an ideal fluid
flow, the energy head remains constant, given by
Q2
E ¼ hþ ¼ const: ð2:38Þ
2gb2 h2
dE dh Q2 dh Q2 db
¼ 2 3 3 2 ¼ 0; ð2:39Þ
dx dx gb h dx gb h dx
or
dh h db
1 F2 F2 ¼ 0: ð2:40Þ
dx b dx
from which the water surface slope at the critical section is (Hager 2010)
1=2
dh h2c d2 b
¼ : ð2:44Þ
dx c 3b dx2 c
Q2 qð x Þ 2
E ðx; hÞ ¼ h þ ¼ hþ : ð2:45Þ
2gh2 bð xÞ2 2gh2
The effect induced by the variation of b with x is better revealed by considering the
unit discharge q = q(x) = Q/b(x). An increase of q produces an increase of the
critical depth and thus a displacement of the E-h function (Fig. 2.10). For a constant
value of the specific energy head Eo across the Venturi flume, critical flow takes
place at the section of minimum width, whereas points 1 and 2 in the curve
corresponding to width B are the up- and downstream sub- and supercritical flow
depths, respectively (Fig. 2.10).
The two basic mechanisms to produce critical flow, namely the (1) contraction of
the channel section and (2) elevation of the bottom profile, can be combined
simultaneously in a transition structure. An example is the inlet to a tunnel diversion
(Fig. 2.11). The inlet structure is used to covey the flow of a river across the
diversion tunnel, while a dam is under construction (Vischer and Hager 1998). The
transition between the river and the circular tunnel can be designed using a rect-
angular channel of gradual decrease in the width and increase in the bottom slope,
to accelerate the river flow from sub- to supercritical tunnel flow, thereby avoiding
backwater effects. Another example is the non-prismatic spillway channel used in
small earth dams (Chow 1959; Castro-Orgaz et al. 2008). The variation of bottom
slope combined with channel width moves the critical flow section away from the
64 2 Energy and Momentum Principles
Q2
H ¼ z b ð xÞ þ h þ ¼ const: ð2:46Þ
2gbð xÞ2 h2
dH dzb dh Q2 dh Q2 db
¼ þ 2 3 3 2 ¼ 0: ð2:47Þ
dx dx dx gb h dx gb h dx
1=2
Rewriting it as function of the Froude number F ¼ Q=ðgb2 h3 Þ produces
dzb dh h db
þ 1 F2 F2 ¼ 0: ð2:48Þ
dx dx b dx
For critical flow F ¼ 1, the location of the critical section must satisfy the relation
dzb hc db
¼ 0: ð2:49Þ
dx b dx
dzb Q2=3 db
1=3 5=3 ¼ 0: ð2:51Þ
dx g b dx
Accounting for the functions zb = zb(x) and b = b(x) in Eq. (2.51) gives a mathe-
matical identity that allows for computing the position of the critical section xc in a
rectangular transition structure under ideal fluid flow.
To illustrate the application of Eq. (2.51), consider an example of Jain (2001)
(Fig. 2.12). A rectangular transition structure ends at a free overfall, involving a
linear width reduction and a parabolic weir profile, given by the equations (di-
mensions in m)
x x x
zb ¼ 2 ; b¼6 : ð2:52Þ
15 15 10
dzb 2 x db 1
¼ 1 ; ¼ ; ð2:53Þ
dx 15 15 dx 10
2 xc 1 Q2=3 xc 5=3
f ð xc Þ ¼ 1 þ 6 ¼ 0: ð2:54Þ
15 15 10 g1=3 10
critical flow section located at the weir crest (dzb/dx = 0) only if the throat (db/
dx = 0) is located there too. This is a practical design approach used for weirs and
flumes for discharge measurement (Montes 1998). In practice, energy dissipation
needs to be accounted for (Castro-Orgaz et al. 2008), as well as the inclusion of
arbitrary cross sections.
Q2
E ¼ hþ : ð2:55Þ
2gA2
Its representation in the E-h plane yields a curve similar to that of Fig. 2.3. The
critical flow condition is in this case
dE Q2 dA
¼1 3 ¼ 0; ð2:56Þ
dh gA dh
Q2
B ¼ 1: ð2:57Þ
gA3
In general, Eq. (2.57) must be solved numerically to compute the critical depth hc
for a given discharge Q and cross-sectional geometry. Inserting Eq. (2.57) into
Eq. (2.55), the minimum (subscript min) specific energy head is
1 Ac
Emin ¼ hc þ : ð2:58Þ
2 Bc
Q2
E ¼ hþa ; ð2:59Þ
2gA2
where a is the Coriolis velocity correction coefficient, given for 1D gradually varied
flow by the expression (Jaeger 1956)
2.2 Energy Principle 67
R R 3
ðu2 þ v2 þ w2 ÞudA u dA
a¼ A
3
A 3 : ð2:60Þ
U A U A
For gradually varied flows, a is of the order of 1.1–1.15 (Chow 1959; Chaudhry
2008), and its consideration is seldom a critical issue (Liggett 1994). However, in
compound channel flows, it can be large, and more importantly, it forces a notable
modification of the behavior of the E = E(h) curve, with two points of minimum
specific energy head (Sturm 2001).
q ¼ ch: ð2:61Þ
Energy conservation for the equivalent steady flow produces (Chow 1959;
Henderson 1966)
c 2 h2 c2
H ¼ h þ Dh þ 2
¼ hþ : ð2:62Þ
2gðh þ DhÞ 2g
Solving Eq. (2.62) for c and assuming that Dh h, given the small wave height,
yields
68 2 Energy and Momentum Principles
" #1=2
2gðh þ DhÞ2
c¼ ðghÞ1=2 : ð2:63Þ
2h þ Dh
The quantity c = (gh)1/2 is the propagation celerity of small gravity waves over still
water. Consider now that the fluid is moving at a constant velocity U (Fig. 2.15a).
This fluid flow produces a displacement of the circular wave pattern generated on
still water. In this case, c is still the wave celerity, i.e., its velocity of displacement
relative to fluid flow. Consider first the case U < c. The absolute displacement
velocity of each wave front is
dx
¼ U c: ð2:64Þ
dt
Given that the wave spreads to both the left and the right, and that c > U, there is
one wave front traveling to the left at speed c–U and another to the right at
U + c (Fig. 2.15b). In still water, the wave patterns are concentric circular fronts
(Fig. 2.16a), but for U < c, there is a displacement of circular waves to the right,
with a wave front propagating to the left (Fig. 2.16b).
Manipulation of Eq. (2.64) yields
dx U
¼Uc¼c 1 ¼ cðF 1Þ; ð2:65Þ
dt c
Fig. 2.16 Wave patterns in a still water, b subcritical flow, c critical flow, d supercritical flow;
adapted from Chow (1959)
U U Fluid velocity
F¼ ¼ ¼ : ð2:66Þ
c ðghÞ1=2 Celerity of small gravity wave
Canterbury, NZ. After war service, he gained the M.Sc. degree from Victoria
University, Wellington, NZ. He then joined in 1952 its staff at the School of
Engineering, where he was particularly involved in the development of the
first axial flow jet boat. After two sabbaticals at the University of Michigan in
1956, and University of Cambridge in 1964, he was appointed as Professor of
hydraulics and Deputy Head of his alma mater. In 1968, he took over the
position of Head of the Civil Engineering Department, University of
Newcastle, NSW. During the following 15 years until retirement, he served
as Dean of Engineering and spent sabbaticals at the University of London in
1974 and at the University of Alberta, Edmonton, CA, in 1977. From 1983,
he was a consultant, with engineering projects in Australia, Southeast Asia,
and New Zealand.
The National Water Committee of the Institution of Engineers, Australia,
decided in 1998 to mark the lifelong achievements of Emeritus Professor
Henderson: During the Henderson Oration, his educational, professional, and
scientific achievements were awarded. The highlights of his professional
career are therein described, including model tests resulting in a shock wave
reduction, vortex generation at intakes, and computer works to investigate
fluid transients. Victor Streeter (1909–2015) has had a great influence on
Henderson’s professional knowledge during his visit to University of
Canterbury in 1952 and during Henderson’s stay at the University of
Michigan. The latter stay marked the initiation of two papers on flow over ski
jumps published in the French journal La Houille Blanche in the early 1960s.
Later, Henderson was involved in the preparation of his famous book Open
Channel Flow, which made him known among hydraulicians and may be
considered his legacy in hydraulics.
ct 1
sin b ¼ ¼ : ð2:67Þ
Ut F
The physical difference between sub- and supercritical flows is illustrated using
an example of Puertas and Sánchez (2001). Consider an obstacle (stone) placed in a
steady and uniform stream. If it is inserted into an initially uniform subcritical
stream (U < c), a gradual water-level variation is produced (Fig. 2.18a) (unsteady
and rapidly varied flow effects are overlooked for the sake of simplicity). The
gradual elevation of the water level is transmitted in the upstream direction as a
smooth wave front. The water level will increase in response to the upstream
discharge received, until the flow starts to spill over the stone, evolving in the long
term to a new steady state (equilibrium condition). Note from this example that the
obstacle influences the conditions of the upstream flow. The obstacle presence is
“informed to the approach flow by small gravity waves.” This information trans-
mission by gravity waves “informs” the upstream flow that there is an obstacle
downstream, so that the flow is able to adapt its condition to that situation.
If the stone is placed into an initially uniform supercritical stream
(U > c) (Fig. 2.18b), there is no possibility of transmitting the information in the
upstream direction, so that the approach flow is not informed about the obstacle
presence. The water then impacts the obstacle and spills over it.1 The present
analysis is qualitative, given that it is in reality an unsteady flow, but highlights the
1
In this introductory example, the possibility of having a hydraulic jump is overlooked. The
hydraulic jump is the way water changes from super- to subcritical flow, to be detailed in the next
section. It is basically a finite amplitude perturbation in steady flow.
2.2 Energy Principle 73
basic difference between sub- and supercritical flows: Subcritical flows are affected
by perturbations produced downstream of a section, whereas supercritical flows are
not affected. Any variation of flow depth or discharge in a channel produces small
gravity wave fronts transmitting information by these perturbations, producing a
transition to a new equilibrium condition.
Q2
BðhÞ ¼ 1; ð2:68Þ
gAðhÞ3
or
AðhÞ3
f ð hÞ ¼ Q 2 g 0: ð2:69Þ
BðhÞ
The problem is, for a given channel shape and discharge Q, to find the value of the
flow depth by which f(h) = 0 in Eq. (2.69). This computation must be generally
conducted using numerical methods. Of the various root-finding methods available
in the literature, the Newton–Raphson method is applied (Hoffman 2001; Jeppson
2011). Consider a generic function f = f(h) sketched in Fig. 2.19. To conduct
iterations of Eq. (2.69), a previous trial value of h is required. Consider the known
value of h at iteration “k”, for which f(hk) is nonzero. The tangent to f(h) at k is
74 2 Energy and Momentum Principles
k
df fk
f 0 hk ¼ ¼ : ð2:70Þ
dh hk hk þ 1
A better approximation to the root of Eq. (2.69) is therefore given by the inter-
section of the tangent line with the h-axis, e.g.,
fk
hk þ 1 ¼ hk : ð2:71Þ
ðdf =dhÞk
that is, if the new iteration value yields an error below a specific tolerance e, usually
taken as 10–6, the actual numerical value is of acceptable quality and therefore taken
as the numerical solution for the critical depth.
For a trapezoidal section (Fig. 2.20), the flow area A and free surface width B are
2.2 Energy Principle 75
A ¼ bh þ zh2 ;
ð2:74Þ
B ¼ b þ 2zh:
A suitable initial value to start iterations is the critical depth of the rectangular
section of the trapezoidal bottom width b
1=3
Q2
hc ¼ : ð2:75Þ
gb2
Q2
3 ðb þ 2zhc Þ ¼ 1; ð2:76Þ
g bhc þ zh2c
Q2
Emin ¼ hc þ 2 ; ð2:78Þ
2g bhc þ zh2c
or
z3 Q2
zEmin zhc gb5
¼ þ
2 : ð2:79Þ
b b 2 1 þ zhbc zhbc
The basic flow variable needed to compute critical flow in a circular cross section of
diameter D is the central angle h
1 hc
h ¼ 2cos 12 : ð2:80Þ
D
The free surface width B and the flow area A are then given by (Jain 2001; Sturm
2001)
h
B ¼ D sin ;
2
ð2:81Þ
D2
A¼ ðh sin hÞ:
8
Q2 1 ðh sin hÞ3
¼ : ð2:82Þ
gD5 512 sinðh=2Þ
The minimum specific energy head follows by substituting Eq. (2.81) into
Eq. (2.58), resulting in
Emin hc 1 ðh sin hÞ
¼ þ : ð2:83Þ
D D 16 sinðh=2Þ
Sub-subcritical flow
Consider the free surface profile in a channel transition consisting of a smooth
hump, to avoid separation losses, located in a prismatic rectangular channel in
which an ideal uniform fluid flow of flow depths up- and downstream of the hump
is established. First, the profiles will be discussed qualitatively using the specific
energy head diagram, and then computations are detailed. Consider ideal fluid flow
toward the hump under subcritical flow with q as the unit discharge, H the total
energy head, and Z the maximum hump elevation (Fig. 2.24). From Eq. (2.14) for
ideal and steady flow,
2.2 Energy Principle 79
Fig. 2.24 Subcritical flow over a hump: a geometry and b specific energy head diagram
H ¼ E1 ¼ E2 þ Z ¼ E3 : ð2:84Þ
E2 ¼ H Z: ð2:85Þ
In this example, it was assumed, as observed on the specific energy head diagram,
that E2 > Emin. The flow depth at point 2 is therefore above hc, and the water
surface forms a depression as it crosses the hump (Figs. 2.24 and 2.25).
Fig. 2.25 Experiment with subcritical flow over a hump [taken from movie Fluid Motion in a
Gravitational Field, by Rouse (1961), IIHR-Hydroscience & Engineering, the University of Iowa]
80 2 Energy and Momentum Principles
Super-supercritical flow
If supercritical flow approaches the obstacle with identical values of H and q as for
F\1, a representation of the flow conditions in the specific energy head diagram
shows that the flow depth over the obstacle increases (Figs. 2.26 and 2.27), again
for E2 > Emin.
The value of Z resulting from equaling E2 and Emin is the maximum possible to
pass the flow rate q with the specific energy head H at the inlet of the hump. It is,
therefore,
Fig. 2.26 Supercritical free surface profile over a hump: a geometry and b representation on the
specific energy head diagram
Fig. 2.27 Experiment with supercritical flow over a hump [taken from movie Fluid Motion in a
Gravitational Field, by Rouse (1961), IIHR-Hydroscience & Engineering, the University of Iowa]
2.2 Energy Principle 81
For this value of Z, the crest flow depth equals the critical depth both for the sub-
and for the supercritical profiles (Fig. 2.28). The computation of the flow profiles is
now detailed.
As the water flows over the obstacle, each section is subject to a different value
of specific energy head imposed by the upstream condition, namely E(x) = H–zb(x).
The problem is to compute the alternate depth corresponding to the flow regime
selected (sub- or supercritical) for the value of E at section x. Equation (2.15) is
written as
3
h E h 2 1
þ ¼ 0: ð2:87Þ
hc hc hc 2
v3 þ av þ b ¼ 0; ð2:88Þ
where
" #
h 1E 1 E 2 1 E 3 27
v¼ ; a¼ ; b¼ 2 þ : ð2:89Þ
hc 3 hc 3 hc 27 hc 2
The two positive solutions to Eq. (2.88) are (Selby 1973; Chanson 2004; Jeppson
2011)
82 2 Energy and Momentum Principles
h E 1 2 C
¼ þ cos ; ð2:90Þ
hc hc 3 3 3
The free surface profile can thus be analytically computed for given zb = zb(x), q,
and H.
For non-rectangular sections, the alternate depths must be determined numeri-
cally. A numerical computation method is presented that is applied to these cases.
The algorithm for the rectangular section is presented, and the numerical solution is
compared with the analytical result. Let the specific energy head at section x be
Eð xÞ ¼ H zb ð xÞ: ð2:93Þ
To determine the flow depth, a root search procedure is applied to the function
q2
f ð xÞ ¼ h þ E ð xÞ: ð2:94Þ
2gh2
f ð xÞ k
hð x Þ k þ 1 ¼ h ð x Þ k ; ð2:95Þ
½df ð xÞ=dhk
df ð xÞ q2
¼ 1 3: ð2:96Þ
dh gh
equation, whereas the specific energy head equals the kinetic energy head for
supercritical flow, e.g.,
8
< E ð xÞ;
2
if 1 gh
q
3 \1
hð xÞk¼0 ¼ h q2 i1=2 ð2:97Þ
: ; if
2
1 gh
q
3 [ 1:
2gEð xÞ
Sub-supercritical flow
An additional possible flow profile involves a change from upstream subcritical
flow to downstream supercritical flow (Fig. 2.31). This profile is obtained by setting
critical flow at the crest and joining the upstream subcritical profile obtained with
Eq. (2.90) with a downstream supercritical profile determined from Eq. (2.91).
Obviously, for this situation to occur, the flow depths up- and downstream from the
obstacle might not be equal. Rather, the two flow depths correspond to the alternate
depths for the specific energy head H = Zmax + Emin (Fig. 2.31).
It is pertinent to mention that for a given upstream subcritical alternate depth h1,
the value of h3 is determined analytically by the alternate depth equation (Fig. 2.31)
q2 q2
H ¼ h1 þ ¼ h 3 þ ; ð2:98Þ
2gh21 2gh23
so that
q2 1 1
h1 h3 ¼ ; ð2:99Þ
2g h23 h21
Fig. 2.31 Transcritical flow profile over a hump: a geometry and b specific energy head diagram
2.2 Energy Principle 85
or
q2 h21 h23 q2 ð h1 h3 Þ ð h1 þ h3 Þ
h1 h3 ¼ ¼ : ð2:100Þ
2g h23 h21 2g h23 h21
q2
h23 h21 ðh1 þ h3 Þ ¼ 0: ð2:101Þ
2g
For given q and h1, Eq. (2.101) is a quadratic equation in h3, which is easily solved.
Consider now how a transcritical flow profile as that of Fig. 2.31 is established
in a channel for given water depths up- and downstream of the obstacle. Consider
upstream subcritical flow (q, H) and a maximum hump elevation Z satisfying
E2 ¼ H Z\Emin : ð2:102Þ
As plotted in the specific energy head diagram (Fig. 2.32), the vertical line labeled
X-X′ does not cut the specific energy head diagram for discharge q. The line
E2 = const. is outside the range of possible physical solutions (Rouse 1938, 1950), so
that the flow rate q cannot be transported over the obstacle with the available specific
energy H E1. The upstream specific energy head must be raised until the specific
energy head on the crest corresponds to the minimum possible Emin for the flow to pass
the obstacle. This is in fact what the water flow does: As the water flow q approaches
the obstacle with the upstream head H, an unsteady flow motion is induced, given the
impossibility to pass the obstacle. The obstacle acts as a perturbation, and this situation
is gradually transmitted to the approach flow by small gravity waves propagating into
the upstream direction at celerity (gh)1/2 relative to the flow. In response to this
feedback, the upstream flow receives the message and begins to increase its upstream
specific energy, establishing a new equilibrium condition. Once the water flows over
the crest at critical flow condition, the upstream energy head is fixed by the new point
1′ as E1′ = Z + Emin > E1. The new head over the obstacle is then E1′. The upstream
specific energy increase from E1 to E1′ is accomplished by a backwater curve of
necessary length to produce an adaptation of flow depths from h1 to h1′ producing the
exact increase needed in the specific energy. This type of curve2 will be described in
Chap. 3, given that frictional effects are relevant.
2
This type of gradually varied flow profile is called M1-type curve. The role of friction is
important, as well as that of the upstream bottom slope So. As will be demonstrated in Chap. 3, a
form equivalent to Eq. (2.8) is dE/dx = So‒Se, with So > Se for the M1 curve, so that E is increased
in the flow direction. A certain length is necessary to increase the specific energy by the quantity
DE needed to pass the obstacle. This specific energy gain is locally lost beyond the obstacle in a
hydraulic jump. The relation describing the energy loss in a hydraulic jump will be presented in the
next section.
86 2 Energy and Momentum Principles
Fig. 2.32 Transcritical ideal flow profile over a hump as result of chocked upstream subcritical
flow: a geometry and b specific energy head diagram
Super-subcritical flow
From the crest, the water flows down the tailwater face of the hump under the
supercritical regime without any energy losses, reaching point 3′ downstream of the
hump. The specific energy head at this point is E1′, which is larger than the value of
the specific energy head E1 far downstream of the obstacle. To return to this flow
condition, the water must lose the specific energy head gained by approaching the
hump, namely E1 − E1′, and change from the super- to the subcritical regime. This
is accomplished by a hydraulic jump, the physical phenomenon producing such
flow transition, accompanied by significant energy losses equal to ΔE = E1′ − E1.
The flow profile over the obstacle is 1-1′-c-3′-3, therefore.
After having described the entire free surface profile, it is observed that the
obstacle “forces” the upstream flow to increase its water depths from h1 to h1′. As
this occurs, namely that the upstream flow conditions are affected, the flow is said to
be “chocked.”
Consider an upstream supercritical flow obeying Eq. (2.102). Again, the
upstream flow must increase its specific energy to pass the obstacle. However, from
gradually varied flow computations to be described in Chap. 3, it is not possible to
gain this specific energy in supercritical flow. To increase the specific energy head
and passing the obstacle, the upstream flow must change from super- to subcritical
flow. Obviously, this transition is accompanied by an energy loss to be supplied by
a backwater curve under subcritical flow. The flow over the obstacle is depicted in
Fig. 2.33. The upstream supercritical flow at section 1 changes to subcritical flow at
section 1′, with the energy loss ΔEj across the hydraulic jump. The energy head at
the crest must be the minimum possible to pass the flow, Emin. Thus, the required
2.2 Energy Principle 87
Fig. 2.33 Transcritical flow profile over a hump resulting in chocked upstream supercritical flow:
a geometry and b specific energy head diagram
energy just upstream of the obstacle is E1′′ = Z + Emin. Therefore, the flow con-
ditions must change from section 1′ to section 1′′ by a gradually varied flow profile
(curve type S1, Chap. 3) of the length necessary to supply the increase in the
specific energy head needed, ΔE, namely (Fig. 2.32)
The upstream supercritical flow is therefore chocked. Beyond the crest, the flow
reaches the toe of the hump under supercritical condition up to section 3′, without any
loss of energy. A gradually varied flow profile in the tailwater portion permits the
adaptation from 3′ to 3. The flow profile over the obstacle is therefore 1-1′-1′′-c-3′.
This type of flow profile is produced if a supercritical flow over an obstacle,
shown in Fig. 2.26, suffers a reduction of discharge; the effect is highlighted in
Fig. 2.34 using the specific energy head diagram, with q2 < q1. The upstream flow
depth h1 is kept constant with a gate. Then, the discharge is reduced, say, by
maneuvers with a pump. The conditions at the upstream section change from point
1 to 1′, which obviously has less specific energy head than the original upstream
flow. This specific energy head is insufficient to pass the obstacle, so it shall be
increased. The only way to increase the specific energy head of a supercritical flow
approaching an obstacle is by jumping to subcritical flow and producing a back-
water effect (Fig. 2.33). Rouse (1961) studied this phenomenon by starting with the
88 2 Energy and Momentum Principles
Fig. 2.34 Discharge reduction in supercritical flow over a hump under constant upstream flow
depth
Fig. 2.35 Experiment with supercritical flow over a hump chocked by discharge reduction [taken
from movie Fluid Motion in a Gravitational Field, by Rouse (1961), IIHR-Hydroscience &
Engineering, the University of Iowa]
Fig. 2.36 Experiment with transcritical flow over a hump [taken from movie Fluid Motion in a
Gravitational Field, by Rouse (1961), IIHR-Hydroscience & Engineering, the University of Iowa]
supercritical profile of Fig. 2.27, then reducing the discharge, and inducing the
propagation of a surge in the upstream direction (Fig. 2.35), which, once reaching
steady state, transforms into a hydraulic jump located in the upstream channel,
leading to a transcritical flow profile over the hump (Fig. 2.36).
2.2 Energy Principle 89
value is reached at the throat for the available value of Eo if the throat width is bmin
as
1=3
3 Q2
Eo Emin ¼ ; ð2:105Þ
2 gb2min
or
3=2
2 3 1=2
bmin ¼ Q gEo : ð2:106Þ
3
For the two cases plotted in Fig. 2.37, bc > bmin. If the up- and downstream flow
conditions are supercritical, the flow passage across the contraction–expansion
produces an elevation–depression of the water surface (Fig. 2.37). If the value of Eo
is kept constant and the value of bc is reduced to bmin, the sub- and supercritical
profiles reach critical flow at the throat (Fig. 2.38).
Fig. 2.40 Transcritical profile in a Venturi flume: a plan and b comparison of test data with theory
1
bð xÞ ¼ bc þ 2 ðx 0:35Þ : ð2:108Þ
8
Consider the measured free surface profile for Q = 0.01 m3/s. The analytical
solution of Eqs. (2.90)–(2.92) was used to generate a theoretical profile, noting that
E = Eo = Emin at all sections, e.g.,
1=3
3 Q2
E ¼ Eo Emin ¼ : ð2:109Þ
2 gb2c
The value of the critical depth at each section is different, however [see
Eq. (2.104)]. That is, upstream from the flume throat, the subcritical flow profile at
section x of the circular arc transition is
hð x Þ Eo 1 2 C
¼ þ cos ; ð2:110Þ
hc ð x Þ hc ð x Þ 3 3 3
where
27 Eo 3
cos C ¼ 1 : ð2:111Þ
4 hc ð x Þ
2.2 Energy Principle 93
Along the 0.05-m-long throat, the water depth is constant and equal to the critical
depth for width bc = 0.12 m. Along the linear expansion, the supercritical flow
profile is
hð x Þ Eo 1 2 C 4p
¼ þ cos þ : ð2:112Þ
h c ð x Þ hc ð x Þ 3 3 3 3
A comparison of the analytical solution and the experimental data reveals fair
agreement, although the water depth along the throat is clearly not a constant.
Deviations of experiments and predictions are mainly due to non-hydrostatic effects
(Hager 2010). Note further the presence of a hydraulic jump in the tailwater, which
is not predicted by the present computations based on constant energy head.
The analytical solution of Eqs. (2.90)–(2.92) was used to compute the free
surface profile of the transition structure shown in Fig. 2.12. A subcritical flow
profile was computed upstream of the critical section using Eq. (2.90), whereas the
supercritical profile downstream from the critical section was determined using
Eq. (2.91). It is thereby necessary to evaluate the specific energy head E(x) and the
critical depth hc(x) at each section. The specific energy head along this transition is
different at each section and given by E(x) = H–zb(x). The total energy head is a
constant to be evaluated using the critical flow section. From Eq. (2.54), it is
located at xc 20.026 m using the Newton–Raphson method. Channel width and
bottom elevation at the critical section are thus bc = 3.997 m and zb = 0.887 m
from Eq. (2.52). The total energy head is then
1=3
3 Q2
H ¼ zb ðxc Þ þ Emin ¼ zb ðxc Þ þ
2 gb2c
1=3 ð2:113Þ
3 302
¼ 0:887 þ 3:573 m:
2 9:81 3:997 2
With the known value of the total head, the application of Eqs. (2.90)–(2.91) is
immediate, using Eq. (2.52) to evaluate b(x) and zb(x) (Fig. 2.41).
The use of the specific energy concept is useful for flows where the energy losses
can be neglected. This amounts to ensure that there are no flow separation zones. If
however, the energy loss is unknown in advance, an alternative method of analysis
must be devised. This method is based on the use of the momentum principle
(Liggett 1994; Montes 1998). Consider for illustrative purposes the hydraulic jump
94 2 Energy and Momentum Principles
Fig. 2.41 Transcritical profile in a flume of parabolic bottom profile with a linear width
contraction (see Fig. 2.12)
(Fig. 2.42a, b). Its internal flow is highly turbulent, with a markedly non-uniform
turbulent velocity profile. The energy loss is unknown, so the energy principle is
unsuited here. Consider steady flow in a horizontal channel where the discharge is
Q and the upstream water depth h1; the problem is to find the tailwater depth h2 of
the hydraulic jump. A control volume to apply the momentum principle is shown in
Fig. 2.42c. Its boundary sections are located outside the zone of rapidly varied flow,
where the pressure is hydrostatic and the velocity distribution close to uniform.
Neglecting the shear forces due to turbulent flow acting on the boundaries of the
control volume, Eq. (2.10) yields
M1 ¼ M2 ; ð2:114Þ
that is, the momentum function M is conserved in the hydraulic jump, with
Q2
M ¼ Ah þ ¼ const: ð2:115Þ
gA
h Q2 h2 q2
M ¼ ðbhÞ þ ¼ b þb ; ð2:116Þ
2 gbh 2 gh
from which arises the definition of specific momentum S (Jaeger 1956; Henderson
1966; Montes 1998)
2.3 Momentum Principle 95
Fig. 2.42 Hydraulic jump: a side view of laboratory test, b top view showing details of
turbulence within the roller, c definition sketch to apply the control volume approach
M h2 q2
S¼ ¼ þ : ð2:117Þ
b 2 gh
The specific momentum S is only a function of flow depth h and unit discharge q, as
the specific energy [see Eq. (2.15)]; it is a quantity conserved in hydraulic jumps.
Equation (2.117) is rewritten as
h2 q2
h S ¼ ¼ const: ð2:118Þ
2 g
This is a cubic in h, plotted in Fig. 2.43. The curve has one asymptote for h = 0,
e.g., the S-axis, and tends to the hydrostatic thrust (1/2)h2 for h ! ∞. As observed,
there is a point of minimum S in Fig. 2.43. This point is determined from the
derivative
dS d h2 q2 q2 q2
¼ þ ¼h 2 ¼h 1 3 ; ð2:119Þ
dh dh 2 gh gh gh
rewritten as
dS dE
¼h ¼ h 1 F2 : ð2:120Þ
dh dh
96 2 Energy and Momentum Principles
This result indicates another property of critical flow (F ¼ 1), namely minimum
specific momentum. The minimum value of S is thus
h2c q2 3
Smin ¼ þ ¼ h2 : ð2:122Þ
2 ghc 2 c
Consider a vertical line plotted in Fig. 2.43, for which the intersection with the
S = S(h) curve yields points 1 and 2. The flow depth h1 < hc and, therefore, the
regime for this point involves supercritical flow. Likewise, for point 2 results
h2 > hc, implying subcritical flow. Points 1 and 2 are determined based on
S = const. and, thus, represent the boundary flow depths of a hydraulic jump. In
general, the two depths of the specific momentum curve obtained by setting
S = const. are called sequent depths or conjugate depths, and shall not be confused
with the alternate depths originating from the specific energy principle.
Equation (2.117) is written in dimensionless form as
1
S 1 h 2 h
¼ þ : ð2:123Þ
h2c 2 hc hc
This universal function is plotted in Fig. 2.44 with Eq. (2.25) for the dimensionless
specific energy head curve. The two curves together reveal that for a hydraulic
jump, which implies S = const., the value of E is not conserved, resulting in an
energy loss. This issue is of relevance to be explored below. For rectangular
channel flow, similar to what was done for the alternate depths as function of E, the
analytical solution for the sequent depths as function of S is from Eq. (2.118), the
solution of the cubic in h/hc
3
h S h
2 2 þ 2 ¼ 0: ð2:124Þ
hc hc hc
The solution for the sequent depths is thus trivially simple for rectangular channel
flow.
The specific energy and momentum concepts are two of the most notable tools
for steady open channel flow analyses. It is widely recognized in the literature that
the specific energy head curve was first developed by Bakhmeteff (1912). The first
diagram is reprinted in Fig. 2.45, corresponding to q = 1 m2/s. Inspecting
Fig. 2.45, it appears evident that the function S = S(h) is plotted superimposed with
the E = E(h) curve. Bakhmeteff denoted S by the letter “h” in his 1912 book.
98 2 Energy and Momentum Principles
Although it appears trivial, it is not, because the two curves of Fig. 2.45 are the
E(h) and S(h) functions, both having been first developed by Bakhmeteff (1912).
This confirms that the origin of the specific momentum concept is simultaneous to
that of the specific energy. Both should be credited to Bakhmeteff (Castro-Orgaz and
Sturm 2018).
Consider now Eq. (2.117) applied between the boundary sections of a hydraulic
jump (Fig. 2.42c)
h21 q2 h2 q2
S¼ þ ¼ 2þ : ð2:128Þ
2 gh1 2 gh2
This equation allows for a relation linking h1 with the unknown value of h2.
Equation (2.128) is rewritten as
2.3 Momentum Principle 99
1 2 q2 1 1
h h2 þ
2
¼ 0; ð2:129Þ
2 1 g h1 h2
or
1 q2 h2 h 1
ðh2 h1 Þðh2 þ h1 Þ ¼ 0: ð2:130Þ
2 g h1 h2
Simplification yields
1 q2
h1 h2 ðh2 þ h1 Þ ¼ 0; ð2:131Þ
2 g
1 1 q2
h1 h22 þ h21 h2 ¼ 0: ð2:132Þ
2 2 g
h2 1 h 1=2 i q2
¼ 1 þ 1 þ 8F21 ; F21 ¼ : ð2:133Þ
h1 2 gh31
This relation is Bélanger’s equation for the sequent depth ratio of a hydraulic jump
in a horizontal, rectangular, and frictionless channel (Bélanger 1849; Montes 1998).
The important point on how momentum and energy principles act comple-
mentary is: The momentum balance applied to a hydraulic jump is a valuable tool
that permits, as shown, to compute the tailwater depth h2 of the hydraulic jump
regardless of the internal flow complexities. Once the momentum principle is
applied, the energy principle can be used to complete the analysis and determine the
hydraulic energy loss, unknown in advance. It is given by
q2 q2 q2 1 1
DE ¼ h1 þ h2 þ ¼ ð h1 h2 Þ þ
2gh21 2gh22 2g h21 h22
ð2:134Þ
q2 h22 h21
¼ ð h1 h2 Þ þ :
2g h21 h22
100 2 Energy and Momentum Principles
q2 1
¼ h1 h2 ðh2 þ h1 Þ; ð2:135Þ
g 2
ð h2 h1 Þ 3
DE ¼ : ð2:136Þ
4h2 h1
The computational process is illustrated graphically with the aid of Fig. 2.46, where
the E(h) and S(h) curves are plotted for the known unit discharge q. With the known
value of h1, the corresponding S value is determined, giving point “a” in Fig. 2.46b.
The vertical line plotted with this value of S yields a cut with S(h) at point “b”,
which is the subcritical tailwater depth h2 of the hydraulic jump. Horizontal lines
plotted from “a” and “b” cut with the E(h) curve, yielding the corresponding
specific energies at both points (points “c” and “d”, respectively). Their difference is
the energy head dissipated by the hydraulic jump. This sequence is conducted
analytically using Eqs. (2.133) and (2.137). Their accuracy is verified in Fig. 2.47,
where the experimental data of Bretz (1988) were used for validation. Note the
small deviation of the latter with predictions due to the neglected boundary shear
effect.
Fig. 2.46 Sequent depth h2 and energy loss DE of a hydraulic jump in a rectangular, horizontal,
and frictionless channel: a hydraulic jump profile and b specific energy and momentum diagrams
2.3 Momentum Principle 101
Q2
M ¼ Ah þ : ð2:138Þ
gA
rewritten as
dM dE
¼A ¼ A 1 F2 : ð2:140Þ
dh dh
Symmetrical trapezoid
For a trapezoidal section of base width b and side slope 1:z, the flow area A is
A ¼ bh þ zh2 ; ð2:141Þ
and the Ah term is, upon decomposing the trapezoid into a rectangle and two
triangles,
h h h2 h3
Ah ¼ bh þ zh2 ¼ b þ z : ð2:142Þ
2 3 2 3
h2 Q2
M¼ ð2zh þ 3bÞ þ : ð2:143Þ
6 g bh þ 12 zh2
To compute the sequent depths for a prescribed value of the momentum function
Mo, a numerical solution is necessary (Jeppson 2011). The Newton–Raphson
formula
fk
hk þ 1 ¼ hk ; ð2:144Þ
ðdf =dhÞk
with
df dM
f ¼ M Mo ; ¼ ; ð2:145Þ
dh dh
yields
ðM Mo Þk
hk þ 1 ¼ hk
k : ð2:146Þ
A 1 F2
2.3 Momentum Principle 103
Circular
Zh Zb
D3 2 D2
A¼ bdg ¼ sin xdx ¼ ðb sin b cos bÞ; ð2:148Þ
2 4
0 0
Zb
1 1
sin2 xdx ¼ b sin b cos b: ð2:149Þ
2 2
0
Zh
Ah ¼ ðh gÞdA; ð2:150Þ
0
Zb
D3
Ah ¼ ðcos x cos bÞsin2 x dx: ð2:151Þ
4
0
With the aid of Eq. (2.149) and the auxiliary primitive function
Zb
1
cos x sin2 xdx ¼ sin3 b; ð2:152Þ
3
0
2.3 Momentum Principle 105
Using Eqs. (2.149) and (2.153), the momentum function of the circular section is
finally (Henderson 1966)
Q2 D3
M ¼ Ah þ ¼ 3 sin b sin3 b 3b cos b
gA 24
ð2:154Þ
4Q2
þ :
gD ðb sin b cos bÞ
2
This function is used to compute the sequent depths of hydraulic jumps as previ-
ously explained for the trapezoidal section if the pipe flow is not pressurized; i.e.,
the subcritical sequent depth does not touch the pipe top.
If the upstream supercritical flow changes to pressurized flow, the hydraulic
jump is said to be “incomplete” (Montes 1998; Jeppson 2011) (Fig. 2.50). For a
transition from upstream supercritical to downstream pressurized flow, the
momentum balance applied to the hydraulic jump control volume produces (Montes
1998; Jeppson 2011)
Q2 D pt Q2
Ah1 þ ¼ At þ At þ : ð2:155Þ
gA1 2 qg gAt
Here, At = pD2/4 is the area of the pipe flowing full and pt the pressure at the pipe
top. Denoting by Mt the momentum function for a gravity pipe flow flowing full
D Q2
Mt ¼ At þ ; ð2:156Þ
2 gAt
pt M1 Mt
¼ ; ð2:157Þ
qg At
Fig. 2.51 Supercritical flow in rectilinear channel with parallel walls involving two steady
sources of perturbations at the inlet
2.3 Momentum Principle 107
ZZZ ZZ
d
qdv þ qðV nÞdA ¼ 0; ð2:158Þ
dt
CV CS
ZZZ ZZ
d
f¼ qVdv þ qVðV nÞdA: ð2:159Þ
dt
CV CS
Here, f is the force vector (resultant of body plus surface forces), V the velocity
vector, n the unit normal to the control surface, q water density, v the volume, and
A the area. Consider in this section steady supercritical flow in a horizontal and
frictionless channel. Further, the parallel-streamlined approach flow has depth h1
and velocity U1 deflected by a wall of angle h (Fig. 2.52a).
Shear forces are neglected, and pressures are assumed to be hydrostatic
(Fig. 2.52b). Due to the wall deflection, the streamlines downstream of the local
perturbation will adjust and become parallel to the tailwater wall direction.
However, this directional change is accompanied by an oblique shock front
crossing the supercritical stream at angle b (Fig. 2.52c).
Consider the control volume sketched in Fig. 2.53 to apply Eqs. (2.158)–
(2.159). Its extreme boundaries are parallel to the shock direction, and a unit width
is considered. The side faces of the control volume are by definition streamlines, so
there is no flow across them. Thus, Eq. (2.158) reduces to
Q1 ¼ Q2 ¼ Q: ð2:160Þ
Fig. 2.52 Oblique shock wave due to abrupt wall deflection: a plan view, b section along
streamline, c photograph of model test
108 2 Energy and Momentum Principles
Fig. 2.53 Application of (- - -) control volume approach to oblique shock wave: a plan and
b section
Equation (2.159) simplifies with fp as the resultant of the pressure forces acting on
the control surface to
1 1
qgh21 qgh22 ¼ qU2n Q qU1n Q: ð2:163Þ
2 2
1 2 1
gh1 þ h1 U12 sin2 b ¼ gh22 þ h2 U22 sin2 ðb hÞ: ð2:164Þ
2 2
Now, Eq. (2.162) is projected in the tangential direction. Note for the control
volume selected that pressure forces acting on each streamline in the tangential
direction are in equilibrium side by side, so there is no net pressure force acting in
this direction. The momentum balance reduces then to
or
h2 1 h 1=2 i U1
¼ 1 þ 8F21 sin2 b 1 ; F1 ¼ : ð2:167Þ
h1 2 ðgh1 Þ1=2
Note that for a finite shock h2/h1 > 1. Only for shocks of infinitesimal height,
namely for h2 h1 results sin b ¼ 1=F1 from Eq. (2.168) as the shallow water
wave solution [see Eq. (2.67)]. These disturbance lines or oblique wave fronts (or
shocks) are sometimes called oblique hydraulic jumps, which is not adequate. The
flow in the downstream portion of these waves is supercritical, so the oblique
standing waves represent transitions from super- to supercritical flow with different
directions of the streamlines. Given that a hydraulic jump is a transition from super-
to subcritical flows, the term “oblique hydraulic jump” should be avoided. Under
chocking conditions, it is possible to have an oblique hydraulic jump, but this is a
particular case. We will therefore denominate these waves as oblique standing or
shock waves. The term “shock wave in supercritical flow” is also a typical
nomenclature in dam hydraulics (Vischer and Hager 1998). The celerity c of the
oblique shock wave is (Ippen 1951)
1 h2 h2 1=2
c ¼ ðgh1 Þ1=2 1þ U1 sin b: ð2:169Þ
2 h1 h1
Given that c equals the velocity component normal to the shock front, the wave is
steady. From Eqs. (2.161) and (2.166), another identity to be satisfied by h2/h1 is
h2 tan b
¼ : ð2:170Þ
h1 tanðb hÞ
tan b 1 h 1=2 i
¼ 1 þ 8F21 sin2 b 1 : ð2:171Þ
tanðb hÞ 2
The ratio of Froude numbers up- and downstream of the shock origin is from
Eq. (2.166) (Jain 2001)
1=2
F2 h1 cos b
¼ : ð2:172Þ
F1 h2 cosðb hÞ
It is of interest to plot in Fig. 2.54 the line defining critical flow in the tailwater
shock portion, e.g.,
1=2
h1
1=2
1 ¼ F1 cos b 1 þ tan2 ðb hÞ ; ð2:173Þ
h2
Using Eq. (2.167) for h2/h1, Eq. (2.174) was numerically solved and the result is
plotted in Fig. 2.54. For a constant value of F1 , b increases with h up to a maximum
value and then decreases. For b = 90° and h = 0, there is no oblique standing wave
but rather a hydraulic jump, given the change from super- to subcritical flow. This
steady wave is not related to the wall deflection, but rather the result of a tailwater
effect. Note from the diagram that for a given value of h two values of b are
possible, one for downstream supercritical flow (F2 [ 1) and the other for sub-
critical flow (F2 \1). The first case is relevant to the practical design of channels
with supercritical flows, given that the wave pattern is only determined by the
upstream flow condition, namely F1 . The angle of the cross-wave and its propa-
gation in the downstream direction by successive reflections at the channel walls
only depends on the upstream conditions. However, for F2 \1, there is a tailwater
effect, propagating a surge against the supercritical flow until a steady wave pattern
2.3 Momentum Principle 111
is formed in conformity with Eq. (2.171). Therefore, this oblique shock with a
transition from super- to subcritical flow depends on a particular combination of up-
and downstream flow conditions. The oblique shock diagram is verified in Fig. 2.55
for supercritical tailwater conditions using the experiments by Ippen and Harleman
(1956).
Consider now supercritical flow in a channel where a wall is deflected by the
angle h, whereas the other wall remains fixed. As previously explained, an oblique
wave front starting at point A crosses the flow at angle b1 reaching the opposite
wall. The flow is subjected thus by a new perturbation at point B, so that the
incident shock is reflected, crossing the flow at a new angle b2. The flow beyond
shock AB becomes parallel to the deflected wall, with a value of F2 \F1 , along
with b2 > b1. This new shock reaches the opposite channel wall at point C and is
reflected with angle b3, with streamlines beyond the shock BC parallel to the
rectilinear wall, and F3 \F2 . The perturbation is further propagated downstream,
tending toward critical flow with each new reflection (Fig. 2.56).
Continuous curved wall with positive deflection
If the wall angle varies continuously as h(x), the continuous curve is divided into
portions of rectilinear walls of discrete deflections Δh to compute the corresponding
Δh across each oblique shock wave (Fig. 2.57). For an upstream value F1 [ 1, any
change Δh due to a flow disturbance occurs along wave fronts, crossing the stream
at angle b. This wave angle is linked to a finite change Δh in the direction of the
tan b tan h
tanðb hÞ ¼ : ð2:175Þ
1 þ tan b tan h
It was already demonstrated that for a positive wall deflection (Fig. 2.53) the
streamlines converge and a discontinuous solution, namely the oblique shock wave,
is formed. In a negative wall deflection, however, streamlines diverge and dis-
continuous solutions are not formed, invalidating the former solutions. Rather, a
continuous depression wave is the correct flow solution (Liggett 1994; Montes
1998). An approximate method applicable to the analysis of negative wall deflec-
tions is now presented. The method is also applicable to positive wall deflections if
2.3 Momentum Principle 113
these are small and shocks are not formed. Consider infinitesimal wave jumps,
namely Δh ! dh, associated with small wall deflections, for which tan b
tan h,
tan h ! h, and h ! dh. Using these simplifications in Eq. (2.177) yields
dh h
¼ : ð2:178Þ
dh sin b cos b
dh U 2
¼ tan b: ð2:180Þ
dh g
For infinitesimal perturbations, the energy loss is negligible and the specific
energy head remains constant. The velocity head is thus
U2
¼ E h: ð2:181Þ
2g
and inserting Eqs. (2.181)–(2.182) into Eq. (2.180) yield the ODE
dh 2ðE hÞh1=2
¼ : ð2:183Þ
dh ð2E 3hÞ1=2
2 3 2 3
1=2
3 1
h ¼ 31=2 tan1 4 1=2 5 tan 4
1
1=2 5 h1 : ð2:185Þ
F 1
2
F 1
2
Equation (2.185) directly links h to F. For the solution produced here, the pertur-
bation lines are no more defining a discontinuous solution. Rather, these are lines of
constant flow depth determined by the boundary angle h(x) at the origin of them on
the wall. An additional relation based on constant specific energy head
E ¼ h½1 þ ð1=2ÞF2 = const. is
h F2 þ 2
¼ 12 : ð2:186Þ
h1 F þ 2
2
h¼ h1 : ð2:187Þ
F
A comparison of Eqs. (2.185) and (2.187) in Fig. 2.58 indicates that the latter is an
accurate approximation of the former if F > 5.
For a continuous convex wall (Fig. 2.59), the streamlines are deflected away
from the original boundary alignment. Due to the negative wall deflection angle
(h < 0), the water depth decreases along the streamlines (Liggett 1994; Montes
2.3 Momentum Principle 115
1998). This case can be analyzed assuming E = const. and thus resorting to
Eq. (2.185). To apply Eq. (2.185) to a positive wall deflection (Fig. 2.57a), the
angle variation must be weak, such that the disturbance lines do not intersect
thereby producing a shock (Liggett 1994). The flow at an abrupt negative wall
deflection (Fig. 2.60a) can be considered as the limit of a convex wall deflection if
the length of the transition curve tends to zero. In this case, the origin of all the
disturbance lines is unique and located at the wall break. The continuous solution is
an expansion fan defined by the lines of constant water depth (Fig. 2.60b) (Liggett
1994). The up- and downstream perturbation angles are
1 1
b1 ¼ sin1 ; b2 ¼ sin1 : ð2:188Þ
F1 F2
For the known values of h1 and U1, the value of F1 ¼ U1 =ðgh1 Þ1=2 is fixed, and the
constant h1 is determined from Eq. (2.185) as
2 3 2 3
1=2
3 1
h1 ¼ 31=2 tan1 4 1=2 5 tan 4
1
1=2 5: ð2:189Þ
F21 1 F21 1
Equation (2.185) is then used to compute F2 for the value of the wall deflection ho,
either numerically or with the aid of Fig. 2.58. Once F2 is known, the flow depth h2
is determined by resorting to Eq. (2.186) and the value of b2 computed by
Eq. (2.188)2. If F1 > 5, it is feasible to use the simpler Eq. (2.187).
Channel contraction
A positive wall deflection produces an oblique shock increasing the water depth,
whereas a negative wall deflection produces a centered depression wave with the
opposite effect. This fact is applied to the design of channel transition structures,
wherein positive and negative waves are superposed to minimize the transmission
of standing waves in the tailwater. The channel contraction (Fig. 2.61) in super-
critical flow is a special case where the wall deflections produce two positive
oblique wave fronts originating at points A and B that intersect and then are
propagated into the tailwater. The length of the channel contraction is selected so
that the wave fronts arrive just at its ends (points C and D). These positive waves
thus interact with the negative waves originating at the same corners, with a the-
oretical cancelation of the wave effects (Ippen and Dawson 1951). This design
approach would work only for the design discharge and is based on assuming that
the interaction of positive and negative waves exactly cancels each other (Vischer
and Hager 1998).
2.4.1 General
A hydraulic control section involves a unique relation between the flow depth h and
the discharge Q and vice versa, e.g., Q = Q(h) or h = h(Q) (Henderson 1966).
A control section is used as boundary condition in water surface profile compu-
tations; thus, their discharge characteristics need to be known in advance (Jain
2001). Further, they are used at selected locations for water discharge measurement
purposes (Bos 1976). Here, the usual control sections in open channel flows are
described, namely uniform flow, critical flow, and the artificial channel control
(Rouse 1950; Jain 2001). The latter is a device inserted into a flow which generates
a head–discharge rating curve. It produces a large perturbation choking the flow,
resulting in a subcritical approach flow that changes to a supercritical flow
downstream of the structure. Examples are weirs and gates. Most artificial channel
controls involve complex flow phenomena as non-hydrostatic pressure, fluid
recirculation, and boundary layers. Here, only an introduction to typical channel
controls is given. Montes (1998) provides an extensive discussion on the fluid flow
fundamentals of head–discharge rating curves of hydraulic structures.
The steady gradually varied flow equation for prismatic channels is [see Eq. (1.110)]
dh So Sf
¼ ; ð2:190Þ
dx 1 F2
where So is the channel bottom slope and Sf the friction slope. Uniform flow is
regained by setting dh/dx = 0 into Eq. (2.190), resulting in
Sf ¼ So : ð2:191Þ
Critical flow is determined from Eq. (2.57), which generates a rating curve Q = Q(hc),
or hc = hc(Q). The frequent location of critical flow sections to start water surface
profile computations is at the maximum elevation of a round-crested weir (Fig. 2.6) or
at the maximum contraction in a non-prismatic channel section (Fig. 2.9). For a
channel with both variable elevation and cross section, the position of the critical flow
section is unknown in advance and must be determined by a singular point analysis
(Figs. 2.11 and 2.12) (Chow 1959; Montes 1998; Hager 2010). Once the critical depth
and its location are determined, the result is used as boundary condition to compute
sub- and supercritical free surface profiles. Note that for supercritical flow computations
in a non-prismatic channel the free surface is bi-dimensional due to standing waves, so
that the one-dimensional gradually varied flow equation breaks down. Another case
where the use of the critical flow theory needs careful attention relates to the flow in
zones of non-hydrostatic flow conditions, as in the vicinity of a free overfall in a
rectangular and horizontal channel, or at the transition from a mild to a steep bottom
slope (Castro-Orgaz and Hager 2017). In the former case, the critical flow section is
located (3–4)hc from the brink section (Fig. 2.62a), and the brink flow depth is hb =
0.715hc (Rouse 1938). The approach flow to the free fall is non-hydrostatic and
supercritical. To conduct gradually varied water surface profile computations, it is
assumed that the critical flow section is located at the brink section, given the small
reach with non-hydrostatic flow (Henderson 1966). Likewise, the flow near a slope
break is non-hydrostatic (Rouse 1938), with the critical flow section shifted upstream to
the small slope reach (Fig. 2.62b). For gradually varied flow computations, the critical
depth is taken at the slope break (Henderson 1966).
Fig. 2.62 Critical flow at a free overfall and b transition from mild to steep slope
2.4 Control Sections 119
Under real fluid flow conditions, there is flow separation around the upstream
corner of the weir (Fig. 2.64a), with the corresponding recirculation of fluid. This
recirculation produces an energy loss and thus a reduction of Cd below its ideal
fluid flow value (2/3)3/2 (Castro-Orgaz and Hager 2017). Other weir types used in
practice are the round-nosed broad-crested weir, the embankment weir, and the
round-crested weir (Fig. 2.64).
If the upstream corner of the broad-crested weir is rounded (Fig. 2.64b), flow
separation is suppressed. The flow will pass above the weir smoothly, with parallel
streamlines, behaving like an inviscid fluid. The exception is the flow close to the
wall, which is influenced by fluid viscosity given the significant reduction in
velocity to comply with the nonslip wall condition. A boundary layer3 starts its
development close to the inlet section, but remains usually thin and rarely reaches
the free surface.
A generalization of Eq. (2.192) is (Montes 1998; Castro-Orgaz and Hager 2017)
1=2
q ¼ Cd gHo3 ; ð2:193Þ
where Ho is the upstream head over the weir. The discharge coefficient of the
round-nosed broad-crested weir is theoretically determined for hydrostatic flow
(Ho/L < 0.33) correcting the critical flow theory by viscous effects resorting to the
boundary layer theory, resulting in dL as the boundary layer displacement thickness
at the weir end section in (Ackers et al. 1978)
3
The boundary layer displacement thickness d* is the virtual shift of the potential flow from a solid
wall due to viscous effects (Montes 1998; White 2009).
120 2 Energy and Momentum Principles
Fig. 2.64 Flow over weirs: a rectangular broad-crested, b round-nosed broad-crested (Ho/L < 0.33),
c embankment, d round-crested
3=2
2 d 3=2
Cd ¼ 1 L ; ð2:194Þ
3 Ho
where
q2
Ho ¼ ho þ w þ : ð2:195Þ
2gðh þ wÞ2
The boundary layer displacement thickness can be estimated with e as the rough-
ness height from (White 2009)
2.4 Control Sections 121
dL e 1=2
¼ 0:001 þ 0:2 : ð2:196Þ
L L
In reality, the flow over a broad-crested weir is surprisingly complex and depends
on the value of Ho/L. For 0.1 < Ho/L < 0.33, the streamlines are parallel over the
weir crest and the pressure distribution is hydrostatic, but viscous effects have to be
accounted for, e.g., by using Eq. (2.195) (Ackers et al. 1978). For 0.33 < Ho/L < 1.5,
the streamlines over the weir are curvilinear and the discharge characteristics are
governed by the non-hydrostatic pressure field. For Ho/L > 1.5, the flow separates
from the crest and the weir behaves like a sharp-crested weir (Montes 1998). For
Ho/L < 0.1, the pressure is non-hydrostatic and viscous effects are relevant, resulting
in an undular weir flow pattern attenuated by friction (Castro-Orgaz and Hager 2017).
Embankment weir
The embankment weir is a broad-crested weir with added slopes up- and down-
stream of the crest (Fig. 2.64c), forming a trapezoid (Hager 2010). The related flow
phenomena are complex, as already discussed, involving action of non-hydrostatic
pressure and bed friction depending on the dimensionless parameter Ho/L (Montes
1998, Hager 2010). From a practical viewpoint, the free flow discharge charac-
teristics are estimated from the empirical equation (Sargison and Percy 2009)
where the effect of the upstream slope angle h is accounted for. The predicted
discharge using Eqs. (2.193)–(2.197) is compared in Fig. 2.65 with data by
Sargison and Percy (2009) relating to trapezoidal weirs with w = 0.25 m and
L = 0.5 m.
The discharge characteristics for free flow are obtained by resorting to the energy
principle and assuming that critical flow takes place on the weir crest [see
Eq. (2.192)], tuning the equations with corrections due to fluid flow phenomena
like non-hydrostaticity or boundary friction. To apply these equations, it is a
requirement that the tailwater flow depth must not affect the control section (Ackers
et al. 1978). An illustrative example on how increased tailwater flow depth affects
weir flow is shown in Fig. 2.66. For a low tailwater depth (Fig. 2.66a), the flow is
critical somewhere on the crest and the discharge characteristics are as for free flow.
A hydraulic jump is formed at the toe of the weir. If the tailwater level is raised, the
hydraulic jump starts moving upstream along the downstream weir face
(Fig. 2.66b), stopping once a steady-state equilibrium in a control volume con-
taining the surge is reached. If the level is further increased (Fig. 2.66c), the
122 2 Energy and Momentum Principles
Fig. 2.65 Computation of rating curves of embankment weir for different upstream slopes
hydraulic jump disappears, and the flow in the vicinity of the slope break will be
non-hydrostatic and near-critical, producing a steady wave train in the tailwater
portion of the structure. For an even larger tailwater flow depth, the structure is
submerged and the flow is subcritical along the entire weir (Fig. 2.66d). The weir
acts as a “control section” for free flow conditions. In the other cases, it behaves like
a transition in bed elevation.
Round-crested weir
The round-crested weir (Fig. 2.64d) is another important type of weir used both as
dam overflow and as a water discharge measurement structure (Montes 1998;
Castro-Orgaz and Hager 2017). Its discharge characteristics are governed by the
non-hydrostatic flow field over the weir (Rouse 1938), with Cd given by
2.4 Control Sections 123
Fig. 2.66 Effect of increasing tailwater depth: a free flow over the crest with hydraulic jump at
weir toe, b hydraulic jump on downstream slope near slope break, c hydraulic jump disappears but
a train of standing waves typical of non-hydrostatic flow is formed, d submerged weir and
subcritical flow
3=2
2
Cd ¼ Co ; ð2:198Þ
3
@2/ @2/
þ 2 ¼ 0; ð2:199Þ
@x2 @z
124 2 Energy and Momentum Principles
with z as the vertical coordinate. For an incompressible fluid, the stream function w
satisfies also the Laplace equation,
@2w @2w
þ 2 ¼ 0: ð2:200Þ
@x2 @z
The lines w = const. are streamlines, and the velocity vector is tangent to them.
The equipotential lines / = const. are normal to the streamlines at intersection
points, forming the flow net of a potential flow (Fig. 2.67) (Rouse 1938; White
2009). While solving the elliptic problem posed by Eqs. (2.199)–(2.200) subject to
appropriate boundary conditions for a given value of Ho, the discharge q emerges as
part of the solution and thus Co. The disadvantage is that complex numerical
methods are required for this task (Montes 1998). Approximate methods to solve
Eqs. (2.199)–(2.200) are available based on the so-called Boussinesq approxima-
tion (Castro-Orgaz and Hager 2017) (see Chap. 11). If the ratio X = Ho/R is small,
a correction to the discharge coefficient of critical (non-hydrostatic) potential flow
over a round-crested weir is with R as the bottom curvature at the weir crest
(Matthew 1963; Castro-Orgaz and Hager 2017)
22 Ho 22
Co ¼ 1 þ ¼ 1þ X. ð2:201Þ
81 R 81
This equation is valid for Ho/R < 0.5 provided that the weir is of a minimum size as
to avoid scale effects originating from viscosity and surface tension (Castro-Orgaz
and Hager 2017). For higher heads, an empirical version of Eq. (2.201) is (Hager
1985)
2.4 Control Sections 125
3X
Co ¼ 1 þ : ð2:202Þ
11 þ 4:5X
q2 q2
E ¼ h1 þ 2
¼ h2 þ : ð2:203Þ
2gh1 2gh22
resulting in
q2 h22 h21
¼ : ð2:205Þ
2g ðh1 þ h2 Þ
The contraction coefficient of sluice gate flow is Cc = h2/a; defining the dis-
charge coefficient Cd as
q
Cd ¼ ; ð2:206Þ
að2gh1 Þ1=2
opening equals the critical depth, the downstream flow turns unstable due to the
presence of standing non-hydrostatic waves typical of near-critical flows (Montes
1998). In this case, there is no uniform flow downstream of the gate, and the
relevance of Cc is obsolete. For larger gate openings, the waves progressively
disappear, leading to a hydrostatic subcritical flow downstream of the gate. For
practical purposes, the gate can be considered to act as a control section for
a/h1 < 0.6, with Cc = 0.6. For larger gate openings, the gate acts as a local
obstruction to the flow.
The force F on the gate is now evaluated by application of the momentum
balance, producing after insertion of Eq. (2.205),
2
h21 q2 h2 q2 1 ðh1 h2 Þ3 1 ðh1 Cc aÞ3
F ¼ qg þ þ ¼ qg ¼ qg :
2 gh1 2 gh2 2 h1 þ h2 2 h1 þ Cc a
ð2:208Þ
This is another example on the complementary usage of the energy and momentum
principles; once the discharge q is determined from the energy principle, the
momentum principle is used to evaluate the force acting on the gate. In dimen-
sionless form, it is
F ð1 Cc a=h1 Þ3
¼ : ð2:209Þ
qg 12 h21 1 þ Cc a=h1
128 2 Energy and Momentum Principles
Consider free flow past a vertical sluice gate in a rectangular channel of equal width
as the gate, with a tailwater flow depth that exactly equals the sequent depth of the
contracted flow depth h2 (Fig. 2.71). A hydraulic jump is therefore formed beyond
the sluice gate. If h3 is lower than the sequent depth h2, a hydraulic jump would still
be formed but further shifted into the tailwater. A water surface profile of super-
critical flow starting at the contracted section should be formed to increase the flow
depth and then equal the actual tailwater level through a hydraulic jump. These
cases are detailed in Chap. 4, given that the surface profile from the vena contracta
section is governed by friction.
It is common in the field monitoring of canal networks to install up- and
downstream of sluice gates ultrasonic sensors to measure h1 and h3. We will form
the resulting system of equations needed to compute the unknowns q and h2 for the
case depicted in Fig. 2.71a. The energy equation applied between the upstream
section and the contracted section is
q2 q2
E ¼ h1 þ 2
¼ h2 þ ; ð2:210Þ
2gh1 2gh22
with the discharge given by Eq. (2.205). The momentum equation applied between
the contracted and the tailwater sections reads
Fig. 2.71 Free flow across sluice gate with rejected hydraulic jump: a definition sketch and
b photograph of model test
2.4 Control Sections 129
h22 q2 h2 q2
S¼ þ ¼ 3þ : ð2:211Þ
2 gh2 2 gh3
It can be rewritten as
1 q2 h2 h3
ð h2 h3 Þ ð h2 þ h3 Þ ¼ ; ð2:212Þ
2 g h3 h2
q2 1
¼ ðh2 þ h3 Þh3 h2 : ð2:213Þ
g 2
Equating q from Eqs. (2.205) and (2.213) results in the quadratic equation for h2
1 1 2 h1 h3 1
h3 h2 þ h3 þ 2h1
2 2
h2 þ h1 h3 ¼ 0:
2
ð2:214Þ
2 2 2 2
Thus, once h2 is computed the discharge q follows using either Eq. (2.205) or
(2.213). This is an illustrative example highlighting how conservation of energy in
the sluice gate problem, and momentum in the hydraulic jump, is used simulta-
neously. In a final step, if needed, the force F on the gate and the energy loss ΔE of
the hydraulic jump are determined by application of the momentum and the energy
balances to the gate and the hydraulic jump, respectively. The use of the energy–
momentum equations in sluice gate problems is known as the EM method.
Consider again the case depicted in Fig. 2.71a. The conjugate depth of the super-
critical contracted flow depth h2 = Cca is
1 h 1=2 i
h02 ¼ Cc a 1 þ 8F21 1 : ð2:215Þ
2
If h2′ = h3, the case of Fig. 2.71a occurs. If h2′ > h3, the jump is further rejected
into the tailwater and the gate operates under free flow conditions, with a backwater
curve starting from the vena contracta section. However, if the gate is operating
under free flow conditions and a tailwater depth is imposed such that h2′ < h3, a
moving hydraulic jump (surge) is shifted upstream, becoming drowned once
reaching the gate section; an alternative flow model is required in this particular
case. The basic assumption, in addition to the hydrostatic pressure distribution, is
that the flow velocity in the roller of the hydraulic jump is small and can be
130 2 Energy and Momentum Principles
q2 q2
E ¼ h1 þ ¼ h þ : ð2:216Þ
2gh21 2gh22
Observe that the kinetic flow energy is given by the velocity in the vena contracta.
Conservation of momentum between the vena contracta and the tailwater section
gives
h2 q2 h2 q2
S¼ þ ¼ 3þ : ð2:217Þ
2 gh2 2 gh3
Equations (2.216)–(2.217) are the EM equations for submerged flow below a sluice
gate (Henderson 1966). In field conditions, ultrasonic sensors are used to measure
h1 and h3. The unknowns are therefore q and h (Fig. 2.72).
Equation (2.216) is rewritten as
q2 1 1
h1 h ¼ ; ð2:218Þ
2g h22 h21
Fig. 2.72 Submerged flow below sluice gate: a definition sketch and b photograph of model test
2.4 Control Sections 131
aY 2 þ bY þ c ¼ 0; ð2:223Þ
1=2
b þ ðb2 4acÞ
Y¼ : ð2:224Þ
2a
From Eq. (2.218), one can write using the definition of Cd given by Eq. (2.206)
h h21 h22 ð h1 þ h2 Þ ð h1 h2 Þ
1 ¼ Cd2 a2 ¼ Cd2 a2 ; ð2:225Þ
h1 h22 h21 h22 h21
132 2 Energy and Momentum Principles
For given values of h1/a and h3/a, Eq. (2.222) is solved for Y and then
Eq. (2.226) for Cd. This task was accomplished and is presented in Fig. 2.73 in the
so-called Henry (1950) diagram. As observed, the theoretical prediction is in fair
agreement with experiments. If a precise water discharge estimation is required,
however, a more advanced model is needed by relaxing some of the starting
assumptions (Castro-Orgaz et al. 2013), as that of zero roller flow velocities or
negligible energy losses across the gate. Note that these computations are to be
conducted only if the tailwater flow depth h3 is larger than the conjugate depth of
h2, given by Eq. (2.215). Computations in Fig. 2.73 were conducted until
Eq. (2.224) produced real solutions. In general, the result to identify a free flow
using Eq. (2.215) is not identical to that derived from Eq. (2.224), due to the special
form of the equations adopted (Jeppson 2011).
gate problem shown in Fig. 2.71. The numerical solution of this problem is for-
mulated defining two functions F1 and F2, for which the roots have to be deter-
mined, namely
q2 q2
F1 ¼ h1 þ h2 þ 0; ð2:227Þ
2gh21 2gh22
2 2
h2 q2 h3 q2
F2 ¼ þ þ 0: ð2:228Þ
2 gh2 2 gh3
or in matrix form
!k !k !
@F1 @F1
F1 @q @h2 Dq
¼ @F2 @F2 : ð2:231Þ
F2 @q @h2 Dh2
With F as the vector of residuals, J the Jacobian matrix, and X the vector of
unknowns,
! ! !
@F1 @F1
F1 @q @h2 q
F¼ ; J¼ @F2 @F2 ; X¼ ; ð2:232Þ
F2 @q @h2 h2
134 2 Energy and Momentum Principles
F ¼ J DX: ð2:233Þ
DX ¼ J1 F; ð2:234Þ
Equation (2.235) must be iteratively applied until the residuals of X are below a
prescribed tolerance. For the present example, the elements of the Jacobian matrix
are
@F1 q q
¼ 2;
@q 2
gh1 gh2
@F1
¼ 1 F22 ;
@h2
ð2:236Þ
@F2 2q 2q
¼ ;
@q gh2 gh3
@F2
¼ h2 1 F22 :
@h2
References
Ackers, P., White, W. R., Perkins, J. A., & Harrison, A. J. M. (1978). Weirs and flumes for flow
measurement. New York: Wiley.
Bakhmeteff, B. A. (1912). O нepaвнoмepнoм движeнии жидкocти в oткpытoм pycлe [Varied
flow of liquids in open channels]. Russia: St Petersburg (in Russian).
Bakhmeteff, B. A. (1932). Hydraulics of open channels. New York: McGraw-Hill.
References 135
Bélanger, J. B. (1849). Notes sur le Cours d’Hydraulique. [Notes on the Course in Hydraulics].
Mém. Ecole Nat. Ponts et Chaussées, Paris, France, session 1849–1850, 222 p (in French).
Blau, E. (1960). Die modellmässige Untersuchung von Venturikanälen verschiedener Grössen und
Form [Model examination of Venturi channels of different sizes and shapes]. Mitteilung 8 der
Forschungsanstalt für Schiffahrt, Wasser-und Grundbau. Berlin: Akademie-Verlag (in
German).
Blau, E. (1963). Der Abfluss und die hydraulische Energieverteilung über einer parabelförmigen
Wehrschwelle [Flow and the hydraulic energy distribution over a parabolic weir]. Mitteilung 7,
5–72. Forschungsanstalt für Schiffahrt, Wasser und Grundbau, Berlin (in German).
Bos, M. G. (1976). Discharge measurement structures. Publication 20. Wageningen, Netherlands:
International Institute for Land Reclamation (ILRI).
Bretz, N. V. (1988). Ressaut hydraulique forcé par seuil [Hydraulic jump forced by a sill].
Communication 2, Laboratoire de Constructions Hydrauliques, EPFL Lausanne, Switzerland
(in French).
Castro-Orgaz, O. (2010). Approximate modeling of 2D curvilinear open channel flows. Journal of
Hydraulic Research, 48(2), 213–224.
Castro-Orgaz, O., Giraldez, J. V., & Ayuso, J. L. (2008). Transcritical flow due to channel
contraction. Journal of Hydraulic Engineering, 134(4), 492–496.
Castro-Orgaz, O., & Hager, W. H. (2017). Non-hydrostatic free surface flows. Advances in
geophysical and environmental mechanics and mathematics (696p). Berlin: Springer. https://
doi.org/10.1007/978-3-319-47971-2.
Castro-Orgaz, O., Mateos, L., & Dey, S. (2013). Revisiting the Energy-Momentum method for
rating vertical sluice gates under submerged flow conditions. Journal of Irrigation and
Drainage Engineering, 139(4), 325–335.
Castro-Orgaz, O., & Sturm, T. W. (2018). Boris A. Bakhmeteff and the development of the
specific energy and momentum concepts. Journal of Hydraulic Engineering, 144(12),
02518004.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). Berlin: Springer.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Hager, W. H. (1985). Critical flow condition in open channel hydraulics. Acta Mechanica, 54(3–
4), 157–179.
Hager, W. H. (2010). Wastewater hydraulics: Theory and practice. Berlin: Springer.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan.
Henry, H. R. (1950). A study of flow from a submerged sluice gate (M.S. thesis). Department of
Mechanics and Hydraulics, State University of Iowa, Iowa City IA.
Hoffman, J. D. (2001). Numerical methods for engineers and scientists (2nd ed.). New York:
Marcel Dekker.
Ippen, A. T. (1951). Mechanics of supercritical flow. Transactions of ASCE, 116, 268–295.
Ippen, A. T., & Dawson, J. H. (1951). Design of channel contractions. Transactions of ASCE, 116,
326–346.
Ippen, A. T., & Harleman, D. R. F. (1956). Verification of theory for oblique standing waves.
Transactions of ASCE, 121, 678–694.
Jaeger, C. (1956). Engineering fluid mechanics. Edinburgh: Blackie and Son.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Jeppson, R. (2011). Open channel flow: Numerical methods and computer applications. Boca
Raton: CRC Press.
Khafagi, A. (1942). Der Venturikanal: Theorie und Anwendung [Venturi flume: Theory and
application]. Versuchsanstalt für Wasserbau, Eidgenössische Technische Hochschule Zürich,
Mitteilung 1. Leemann, Zürich, Switzerland (in German).
Liggett, J. A. (1994). Fluid mechanics. New York: McGraw-Hill.
Matthew, G. D. (1963). On the influence of curvature, surface tension and viscosity on flow over
round-crested weirs. Proceedings of ICE, 25, 511–524; 28, 557–569.
136 2 Energy and Momentum Principles
Montes, J. S. (1998). Hydraulics of open channel flow. Reston VA: ASCE Press.
Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. (2007). Numerical recipes:
The art of scientific computing (3rd ed.). Cambridge: Cambridge University Press.
Puertas, J., & Sánchez, M. (2001). Apuntes de Hidráulica de canales [Open channel hydraulics
lecture notes]. Civil Engineering School, University da Coruña, Spain (in Spanish).
Roth, A., & Hager, W. H. (1999). Underflow of standard sluice gate. Experiments in Fluids, 27(4),
339–350.
Rouse, H. (1938). Fluid mechanics for hydraulic engineers. New York: McGraw-Hill.
Rouse, H. (1950). Engineering hydraulics. New York: Wiley.
Rouse, H. (1961). Fluid motion in a gravitational field: IIHR Movies, University of Iowa.
Sargison, J. E., & Percy, A. (2009). Hydraulics of broad-crested weirs with varying side slopes.
Journal of Irrigation and Drainage Engineering, 135(1), 115–118.
Selby, S. M. (1973). Standard mathematical tables. Cleveland: CRC.
Sturm, T. W. (2001). Open channel hydraulics. New York: McGraw-Hill.
Subramanya, K. (1986). Flow in open channels. New Delhi: Tata McGraw-Hill.
Toman, E. M., Skaugset, A. E., & Simmons, A. N. (2014). Calculating discharge from culverts
under inlet control using stage at the inlet. Journal of Irrigation and Drainage Engineering,
140(2), 06013003.
Vischer, D. L., & Hager, W. H. (1998). Dam hydraulics. New York: Wiley.
White, F. M. (2009). Fluid mechanics. New York: McGraw-Hill.
Yen, B. C. (1973). Open-channel flow equations revisited. Journal of the Engineering Mechanics
Division, ASCE, 99(EM5), 979–1009.
Chapter 3
Computation of Steady
Gradually-Varied Flows
3.1 Introduction
The momentum principle yields for a prismatic channel the system [Eq. (1.100)],
dM
¼ gA So Sf ; ð3:1Þ
dx
Q2
M¼ þ Ah: ð3:2Þ
gA
Here, M is the momentum function, Sf the friction slope, Q the discharge, A the flow
area, x the streamwise coordinate, g the gravity acceleration, h the depth of the
section centroid below the free surface, and So the bottom slope. The
gradually-varied flow equation resulting from Eqs. (3.1)–(3.2) is [Eq. (1.110)],
so
dh So qgRh So Sf
¼ ¼ ; ð3:3Þ
dx 1 B Q23
gA
1 F2
where so is the boundary shear stress, Rh = A/p the hydraulic radius, p the wetted
perimeter, B the free surface width, h the flow depth, q the water density, and F the
Froude number. Equation (3.3) was first presented by Bélanger (1828).
The energy principle yields, with E as the specific energy and Se the gradient of
dissipated energy (Montes 1998; Chanson 2004) the system
dE
¼ So Se ; ð3:4Þ
dx
Q2
E ¼ hþ : ð3:5Þ
2gA2
Equation (3.3) is obtained from Eqs. (3.4)–(3.5) if the gradient of dissipated energy Se
equals the friction slope Sf (Yen 1973, 1991, 2002; Montes 1998). Note that the
velocity distribution coefficients are neglected in the systems (3.1)–(3.2) and (3.4)–(3.5).
Given that the energy (Coriolis) and momentum (Boussinesq) velocity correction
coefficients are different theoretically, gradually-varied flows are defined by a unique
ODE [Eq. (3.3)] if and only if the velocity correction coefficients are assumed to be
unity and the friction slope equals the gradient of dissipated energy. Both slopes are
different from a theoretical point of view (Yen 1973, 2002): the friction slope is a
measure of the shear forces acting on the channel boundaries, whereas the gradient of
dissipated energy is a measure of all the losses accumulated by the entire mass of fluid.
For gradually-varied flows in a prismatic channel, both are equal once the velocity
distribution is assumed to be uniform. At abrupt variations of the channel
cross-section, e.g., expansions or contractions, local energy losses due to eddies
become significant. These additional losses have to be introduced empirically into
gradually-varied flow computations by using local energy-loss coefficients (Montes
1998). In this chapter, we limit the analysis to flow profiles along prismatic reaches.
3.3.1 Definition
Uniform flow is the simplest case of non-uniform flow, obtained by setting the
variation of depth with distance equal to zero in Eq. (3.3), that is,
3.3 Uniform Flow 139
dh
¼ 0 ) So Sf ¼ 0 ) so ¼ qgRh So : ð3:6Þ
dx
The constant flow depth satisfying Eq. (3.6) is defined as the normal or uniform
flow depth hN. In uniform flow, the friction slope is identical to the channel bottom
slope. To compute the uniform flow depth for a given So resorting to Eq. (3.6), a
parametrization of the bed-shear stress so as function of the flow variables is
required. This task is accomplished below.
Rouse (1965) parametrized the boundary shear stress in open channel flows by
defining the general relation
so
¼ HðR; F; e; C; N; UÞ: ð3:7Þ
qU 2
Here, R ¼ UD=m is the Reynolds number, U the mean velocity, D the hydraulic
diameter, m the kinematic viscosity, F the Froude number, e = ks/D the relative
roughness, ks the equivalent wall roughness height, C the effect of the
cross-sectional shape, N the effect of non-uniform flow, and U the unsteadiness
effect. While Eq. (3.7) is a general statement from dimensional analysis, it is dif-
ficult to find particular expressions in practice, given the nonlinear interactions of
the various hydraulic numbers (Yen 2002). For the specific case of steady uniform
flow, Eq. (3.7) reduces to (Yen 2002)
so f
¼ HðR; eÞ ¼ ; ð3:8Þ
qU 2 8
f
qU 2 ¼ qgRh Sf ; ð3:9Þ
8
1 U2 1 U2
Sf ¼ f ¼f : ð3:10Þ
4Rh 2g D 2g
140 3 Computation of Steady Gradually-Varied Flows
Note that in open channel flows D = 4Rh. The problem is thus reduced to find a
suitable predictor for H(R, e) in Eq. (3.8). Open channel flow is generally turbulent
(R > 2000) (White 1991, 2003), and attention is therefore limited to this regime.
A Colebrook-White type formula suitable for open channel flows is (Yen 2002)
4e c
f 1=2 ¼ a log10 þ ; ð3:11Þ
b 4Rf 1=2
where the empirical coefficients (a, b, c) adopt different values, depending on the
literature source [see Yen (2002) for a detailed discussion]. Here, the values pro-
posed by Henderson (1966) (a = 2, b = 12, c = 2.5) are used. Equation (3.11) is
plotted in Fig. 3.1 using these coefficients, by numerically solving it for given
values of R and e.
The limiting form of Eq. (3.11) for turbulent smooth flows (e = 0) is
1=2 2:5
f ¼ 2 log10 ; ð3:12Þ
4Rf 1=2
whereas for turbulent rough flows (R ! ∞) one gets the explicit equation
2
1 1
f ¼ log10 e : ð3:13Þ
4 3
In hydraulic smooth flows, the wall roughness elements are submerged into the
laminar viscous sub-layer, and the flow resistance is fully determined by R (White
1991, 2003). In hydraulic rough flows, the roughness elements emerge above the
thin viscous sub-layer, and the flow resistance is a form drag, fully determined by e
(White 1991, 2003). Between these two limits, the flow is transitional, and the flow
resistance is a function of both R and e. Many practical open channel flow problems
are likely to be fully rough, resorting to Eq. (3.13) and greatly simplifying com-
putations, therefore. However, smooth flows are also usual in open channel laboratory
facilities. Thus, it is equally important to have a simple predictor for f in these flows.
As noted from Fig. 3.1, a good explicit approximation to Eq. (3.12) is (Yen
2002; Fenton 2010)
1 1:95 2
f ¼ log10 : ð3:14Þ
4 R0:9
By simply combining the limiting forms for rough and smooth flows, given by
Eqs. (3.13) and (3.14), the following explicit formula is obtained (Yen 2002,
Fenton 2010)
1 e 1:95 2
f ¼ log10 þ : ð3:15Þ
4 3 R0:9
This equation is plotted in Fig. 3.1 and compared to Eq. (3.11), resulting in an
excellent match. Applying Eq. (3.15) in combination with Eq. (3.10) provides a
simple and physically sound method to compute flow resistance in open channels.
Usual values of the roughness height are given in Table 3.1.
Another widely used flow resistance relation for turbulent rough open channel
flows is the Gauckler–Manning–Strickler (GMS) equation (Sturm 2001; Yen 2002)
n2 U 2
Sf ¼ 4=3
; ðSIÞ ð3:16Þ
Rh
f gn2
¼ 1=3 ; ð3:17Þ
8 R
h
ks1=6
n¼ : ðSIÞ ð3:18Þ
21:1
Fenton (2010) took Eq. (3.18) as reference and conducted the corresponding
comparison, concluding that agreement between Eqs. (3.15)–(3.17) and Eq. (3.18)
results if the coefficient 21.1 in Eq. (3.18) is reduced by 10–20%. Chaudhry (2008)
quoted for computing n as function of ks the formula
ks1=6
n¼ ; ðSIÞ ð3:19Þ
8:25g1=2
This is compared in Fig. 3.2 with Eq. (3.13) for the fully rough flow, rewritten as
2
gn2 1 1 1
1=3
¼ f ¼ log10 e : ð3:21Þ
Rh 8 32 3
3.3 Uniform Flow 143
n0 ¼ ng1=2 : ð3:22Þ
n0 2 U 2
Sf ¼ 4=3
: ð3:23Þ
gRh
The unit of n′ is L1/6 in SI, which is reasonable for the measure of a roughness; it is
only related in this way to the dimension length. Note that Eq. (3.22) is implicit in
the empirical Eq. (3.19), which relates n′ to k1/6 s , with ks as a physical-based
quantity. It permits to write Eq. (3.23) as (Chaudhry 2008)
Equation (3.24) is a modified Manning equation suitable for river flows simply
replacing ks by a mean particle diameter of the bed surface layer.
In this section, the computation of the normal depth will be presented using
Manning’s equation with a constant value of n. Similar computations are accom-
plished by adopting the Darcy–Weisbach equation or a variable n with flow depth
resorting to Eq. (3.17). Inserting the continuity equation Q = UA into Eq. (3.16)
produces, with Sf = So,
S1=2 2=3
Q¼ o
ARh : ð3:25Þ
n
144 3 Computation of Steady Gradually-Varied Flows
nQ 2=3
1=2
¼ ARh ; ð3:26Þ
So
nQ
f ðhÞ ¼ AðhÞ½Rh ðhÞ2=3 1=2
0: ð3:27Þ
So
Therefore, for a given channel shape and values of Q, n and So, the flow depth by
which f(h) = 0 in Eq. (3.27) is the uniform flow depth hN. The computation must be
conducted using numerical methods. Given a known value of h at iteration “k” for
which f(hk) 6¼ 0, a better approximation to the root of Eq. (3.27) is given by the
Newton–Raphson iterative formula (Hoffman 2001; Jeppson 2011)
fk
hk þ 1 ¼ hk : ð3:28Þ
ðdf =dhÞk
Equation (3.28) is applied recursively until the difference between two successive
approximations is down to a prescribed tolerance e 10−6, e.g.,
3.3 Uniform Flow 145
kþ1
h hk
6
hk
10 : ð3:31Þ
A ¼ bh þ zh2 ;
B ¼ b þ 2zh;
p ¼ b þ 2hð1 þ z2 Þ1=2 ; ð3:32Þ
A
Rh ¼ :
p
The wetted perimeter derivative needed to apply Eq. (3.30) for the trapezoidal
section is
dp
¼ 2ð1 þ z2 Þ1=2 : ð3:34Þ
dh
Fig. 3.3 Uniform flow in trapezoidal channel a streamwise profile, b transverse section
3.4 Flow Profiles in Prismatic Channels 147
S1=2 2=3
Q o
ARh ¼ 0 ) hðxÞ hN ) So ¼ Sf : ð3:35Þ
n
Given that the channel is prismatic, the normal depth is represented by a straight
line parallel to the bed along the channel, called normal depth line (NDL).
Further, the critical depth hc is determined by the implicit equation (Chap. 2)
A3
Q2 g ¼ 0 ) hðxÞ hc ) F ¼ 1: ð3:36Þ
B
Similar arguments reveal that the critical depth is also a constant along a prismatic
channel. If we plot the critical depth in the channel, it is represented by a straight
line parallel to the bed, referred to as the critical depth line (CDL).
As an example, the NDL and CDL are plotted in Fig. 3.5 for a channel with
constant Q for different values of the bottom slope. As the critical depth is inde-
pendent of the bottom slope, in the absence of cross-section variations, it remains
unchanged. Note that for the cases (a) and (b), the normal depth is above the critical
depth, but for case (c), the bottom slope is large, so that the normal depth is below
the critical depth. This suggests that there is a particular value of the bottom slope
for which the normal depth exactly equals the critical depth. This value is called
critical slope Sc; it is determined as follows. The normal depth equation is
n2 Q 2
So ¼ 4=3
: ð3:37Þ
A2 Rh
Note that the critical slope is determined by n, the cross-sectional geometry, and the
flow depth hN = hc. As hc is a function of Q, Sc is also variable with it. However,
the critical slope Sc is independent of the actual bottom slope. The following
classification is therefore feasible:
1
In the present context, a steep slope implies that the normal depth is below the critical depth.
However, a steep slope as used here shall be “mild” physically, that is, 1 + S2o 1. Otherwise
slope corrections are necessary in the GVF Eq. (3.3) [see Chap. 1, Eq. (1.151)]. A steep slope in
hydraulic structures implies that 1 + S2o > 1, to be discussed at the end of this chapter.
150 3 Computation of Steady Gradually-Varied Flows
dh So Sf þ
¼ ¼ ¼ þ: ð3:39Þ
dx 1 F2 þ
The M1 type flow curve thus increases in the flow direction. Following the same
methodology, all the possible M-type flow profiles are summarized in Table 3.3.
The column dE/dx = So − Sf indicates the profiles with either an increase or
decrease of the specific energy.
Fig. 3.8 Flow zones in steep sloping channels [adapted from Puertas and Sánchez (2001)]
For the steep slope channels, the flow zones are presented in Fig. 3.8 following the
same methodology. As example, the supercritical (F > 1) flow in zone 2 (hc > h > hN)
has Sf < So, resulting in dh/dx < 0, that is, the S2 curve is decreasing in the flow
direction. Following the same procedure, a classification of the remaining S-type
flow profiles entails no difficulty; all the possible profiles are summarized in
Table 3.3, including the critical, horizontal, and adverse flow profiles.
To plot the water surface profile, it is necessary not only to know the sign of
dh/dx in the flow zone, but also the tendency of the water surface profile toward the
reach boundaries (Jain 2001; Chaudhry 2008). The following trends prevail as the
flow depth approaches the critical depth, normal depth, or becomes very large:
dh
h ! hc ) ! 1 if So 6¼ Sf ;
dx
dh
h ! hN ) ! 0 if F 6¼ 1; ð3:40Þ
dx
dh
h!1) ! So :
dx
As the flow approaches the critical depth, the water surface slope becomes vertical.
It simply implies that the critical depth is not a mathematical solution of GVF in
prismatic channels, whereas it is for non-prismatic channels of variable slope, to be
discussed in Chap. 4. As the flow approaches the normal depth, the profile tends to
the NDL, e.g., the normal depth is an asymptotic value only. As the flow depth
becomes very large, the velocity tends to zero, as do both Sf and F. It implies
that the flow depth gradient equals So, i.e., the flow surface is horizontal,
d(h + zb)/dx = 0 (remember that So = –dzb/dx).
152 3 Computation of Steady Gradually-Varied Flows
Consider now the M1 type water surface profile. An ascending free surface
profile in zone 1 must approach to the left the uniform flow depth, and to the right
produce large water depths. From Eq. (3.40), the flow profile approaches in the
upstream direction asymptotically the uniform flow depth, and to the right it has a
horizontal asymptote (Fig. 3.9). For the M2 water surface profile, there is a trend to
uniform flow to the left, and to critical depth to the right, given that dh/dx < 0. The
corresponding flow profile is sketched in Fig. 3.10.
For the S2 curve in a steep slope, the flow starts close to the critical depth and
gradually approaches the uniform flow (Fig. 3.11).
Following the same methodology and combining the cases of Table 3.3 with
Eqs. (3.40), all possible free surface profiles are presented in Fig. 3.12. Several
real-life cases are presented in Fig. 3.13 following Chow (1959).
3.4 Flow Profiles in Prismatic Channels 153
Fig. 3.12 Flow profiles in a mild, b steep, c horizontal, d critical, and e adverse slopes
154 3 Computation of Steady Gradually-Varied Flows
Fig. 3.13 Real-life cases of water surface profiles [adapted from Chow (1959)]
3.4 Flow Profiles in Prismatic Channels 155
Danny Lee Fread was born on 17 July, 1938, at Tuscola IL, passing away
aged 70 years on 5 February, 2009, at Huntington PA, USA. He worked a
plethora of jobs to pay his way through his Liberal Arts and Engineering
degrees from Carthage College, Kenosha WI, and the University of
Missouri-Rolla, Rolla MO. After working for Texaco, he returned to his
Alma Mater to continue his education, receiving the Ph.D. degree in civil
engineering in 1971. He joined the Office of Hydrology at the National
Weather Service then in Silver Spring MD as a research hydrologist. Inspired
by the tragedy of the failure of Grand Teton Dam in 1976, his research
focused on developing computer models to forecast the flow of flooding
rivers and dam failures. His computer models were used around the globe.
Fread received national awards for his work, including the Department of
Commerce Gold Medal, the Huber Research Prize from the American Society
of Civil Engineers ASCE, its 1976 J. C. Stevens Award, and the Association
of State Dam Safety Officials National Award of Merit. He also was a Fellow
of the American Meteorological Society. He ended his career as Director of
the Office of Hydrology. Following his retirement, he moved with his wife to
Pennsylvania to be near their daughter and family.
The 1973 ASCE paper presents a conceptual model to alleviate flood
damages due to overtopping failures of future small earthfill dams including
the erosion pattern. The potential reduction in the reservoir release due to the
proposed erosion-retarding layer is also investigated. A method to determine
the optimum layer location is provided to minimize the maximum possible
reservoir release due to a gradually-breached earth dam. The transient
reservoir flow is simulated by a numerical model based on the solution of the
one-dimensional Saint-Venant equations, solved by the method of charac-
teristics subjected to appropriate boundary conditions. The numerical simu-
lation provides the reduction in release discharge in terms of various
parameters.
156 3 Computation of Steady Gradually-Varied Flows
In general, a real channel presents reaches of different lengths, slopes, sections, and
roughness. The actual free surface profile will be a composite surface assembling
some of the elementary profiles depicted in Fig. 3.12. Before computing profiles
numerically, the sketching process is as follows (Chow 1959; Jain 2001; Chaudhry
2008):
1. Plot the bottom profile (use distorted scale in the vertical) and assign roughness
values to each reach.
2. Compute hN for each channel reach using Eq. (3.35) and plot the NDL.
3. Compute hc for each channel reach using Eq. (3.36) and plot the CDL.
4. Locate all possible control sections, namely a critical depth control (CDC),
normal depth control (NDC), and artificial channel control (ACC) (e.g., a weir
or a gate). In a CDC, the flow is subcritical upstream and supercritical down-
stream. The probable position of a NDC is upstream of the reach in subcritical
flow and downstream of the reach in supercritical flow. A NDC is an asymptotic
state. In an ACC, the flow is subcritical upstream of it and supercritical
downstream.
5. Starting at each control section, draw sub- and supercritical profiles in the
up- and downstream directions, respectively. Check possible interactions of
control sections.
6. If the flow is supercritical upstream of the reach and subcritical downstream, a
hydraulic jump may form within the reach. The jump may also not be formed,
resulting in fully sub- or supercritical flow along the reach. The position of the
jump can be plotted tentatively, though its formation will be discussed in depth
in Chap. 4.
As an example, two illustrative cases presented by Subramanya (1986) are
plotted in Fig. 3.14. In case (a), the three slopes of the channel are mild. Note that
the values of the NDL for each reach are above the CDL. The gate on the reach c–e
is an ACC that produces subcritical flow upstream and supercritical flow down-
stream. At the free overfall (section e), the critical depth is attained at a distance of
three to four times hc from the brink section (Rouse 1938). Given that this distance
is small, the critical depth control section is assumed to be established just at the
brink in GVF computations. Thus, the brink section is a control section interacting
with the vena contracta section (point d), which is the control section induced by the
gate. From the vena contracta, a supercritical M3 profile develops, whereas the
brink section will force a subcritical M2 profile. To have compatibility of both
profiles, a hydraulic jump is formed within the reach. Cases where the gate is
drowned or the jump rejected out of the reach will be discussed in Chap. 4. At point
c, the flow is subcritical as forced by the gate. Starting at this point, a M1 subcritical
profile is formed in the reach b–c. If this reach is long, as assumed here, the normal
depth will be asymptotically approached and established at point b. From point b,
an M2 profile is formed up to the upstream section a.
3.4 Flow Profiles in Prismatic Channels 157
Fig. 3.14 Examples of GVF profiles with various controls [adapted from Subramanya (1986)]
In case (b), a mild slope is followed by two steep slopes. Note again the values
of the NDL as compared with the CDL. In the absence of any other control, it is
assumed that the normal depth (NDC) is established upstream in reach a–b, and
downstream in reach b–c. As the flow in the mild slope reach is subcritical, and
supercritical on the steep slope, a CDC is established, in this case just at the slope
break (point b). The flow profile upstream of b is therefore an M2 profile, whereas
downstream of it, an S2 profile develops. At point c, a NDC was assumed, so that
an S3 profile is formed within reach c–d.
The sketching methodology for the free surface profile previously explained is
based on the assumption of known discharge. However, sometimes, the discharge is
unknown and must be determined as part of the solution itself. The case of a
channel releasing water from a reservoir is an example (Fig. 3.15). Consider a
constant slope channel of rectangular cross-section of width b. The basic known
magnitude is the water surface elevation at the reservoir. If the elevation of the
channel invert intake is known, then the water depth at the channel inlet H is
available. In the forthcoming analysis, the velocity head in the reservoir and the
intake entrance losses are neglected.
The problem is thus to determine the discharge Q released from the reservoir by
the channel. As the discharge is unknown, it is not possible to compute hN and hc,
158 3 Computation of Steady Gradually-Varied Flows
which are needed to determine the type of control section of the flow profile. Thus,
the solution must be conducted by iteration. Assume first that the channel is steep
(Fig. 3.15a). In this case, the critical depth is given by
2
hc ¼ H; ð3:41Þ
3
which is established at the channel inlet, with the released discharge given by
3=2
2 3 1=2
Q¼b gH : ð3:42Þ
3
Now, the steep slope assumption shall be checked. Therefore, the critical slope Sc is
computed by resorting to Eq. (3.38), and compared to the channel slope So. If
So > Sc, then the initial assumption is correct, the channel is thus steep and com-
putations are finished. The normal depth is not needed to compute Q, but it can be
determined from Manning’s equation to have the complete solution to the flow
conditions. If So < Sc, the channel is not steep, and computations need to be
reconsidered. For a mild slope channel (Fig. 3.15b), uniform flow is established
assuming that the channel is long. In this case, a normal depth control governs the
channel intake, resulting in the energy equation
Fig. 3.15 Discharge released from a reservoir for a mild and b steep slope channels
3.4 Flow Profiles in Prismatic Channels 159
Q2
H ¼ hN þ : ð3:43Þ
2gb2 h2N
Resorting to Eq. (3.38), the critical slope Sc can be updated for the correct discharge
Q. This solution process was implemented in the file Lake_discharge.xls available
from Chap. 12. As application example, consider a rectangular channel of b = 4 m,
n = 0.013 sm−1/3 releasing water from a reservoir with H = 1 m. For a slope
So = 0.01, the program yields hc = (2/3) m, Q = 6.82 m3/s, hN = 0.438 m and
Sc = 0.00278. The channel is therefore steep and the control section is a CDC at the
channel inlet (Fig. 3.15a). If the slope is So = 0.0001, a comparison with Sc shows
that the steep slope assumption is incorrect, and computations are conducted again
assuming mild slope with a NDC at the inlet. The program then yields the output
hc = 0.322 m, Q = 2.289 m3/s, hN = 0.983 m and Sc = 0.00295. Given that
0.0001 < 0.00295, the mild slope assumption is correct.
The GVF equation is a first-order ODE describing the variation of h(x), written as
dh So Sf
¼ ¼ f ðhÞ: ð3:47Þ
dx 1 F2
160 3 Computation of Steady Gradually-Varied Flows
which is rewritten as
dh
¼ f ðx; hÞ: ð3:49Þ
dx
Zx2
E2 E1 ¼ ðSo Se Þdx; ð3:50Þ
x1
Zx2
E2 ¼ E1 þ So ðx2 x1 Þ Se dx: ð3:51Þ
x1
To apply Eq. (3.51), the integral of the gradient of dissipated energy (or energy
slope) is numerically approximated. Techniques to solve Eq. (3.49) or its integral
form, Eq. (3.50), will be presented in the next sections.
3.5 Computation of Steady Flow Profiles 161
dh
¼ f ðx; hÞ;
dx ð3:52Þ
hðx ¼ xo Þ ¼ ho :
Fig. 3.17 Control and gauging sections in transition from mild to steep slope
hðx ¼ xo Þ ¼ hc e; ð3:53Þ
where e is a small number. The inverse transition from super- to subcritical flow
occurs in the form of a hydraulic jump, and thus is not a control section.
From this discussion follows that the “usual” position of the control section is at
the downstream end in a subcritical reach. Likewise, the usual position of a control
section in a supercritical flow reach is at its upstream end. Accordingly, the GVF
computational rule of determining the profiles is in the upstream direction for
subcritical flow and in the downstream direction for supercritical flows. This is in
agreement with the physical fact that perturbations travel both up- and downstream
for subcritical flow, thus being influenced by what occurs at the end of the reach,
whereas supercritical flows are only influenced by perturbations induced upstream
of the reach. Subcritical flows are thus subject to a downstream control, whereas
supercritical flows to an upstream control. This will be further discussed once
unsteady flow is introduced in Chap. 5.
Consider now the setup of Fig. 3.17 in a laboratory facility. An experimentalist
may take a measurement at the gauging section xo and then get a known ho for Q,
which is different from the standard control section at the slope break. This section
could be used to integrate Eq. (3.49) in the up- and downstream directions, by
which a portion of the subcritical profile would be computed in the downstream
3.5 Computation of Steady Flow Profiles 163
hðx ¼ xo Þ ¼ hN e: ð3:54Þ
Q2 Q2
So S o
dh C 2 A2 Rh C 2 b2 h2 Rh
¼ 2
¼ : ð3:55Þ
dx Q Q2
1 3B 1 2 3
gA gb h
Assuming that the channel is very wide, the hydraulic radius is approximated by the
flow depth, resulting in the GVF equation
164 3 Computation of Steady Gradually-Varied Flows
Q2 Q2 h3
So 1 1 N3
dh 2 2 3
C b h ¼S 2 2
So C b h 3
h :
¼ o ¼ So ð3:56Þ
dx Q2 Q2 h3
1 2 3 1 2 3 1 c3
gb h gb h h
In Eq. (3.56), hN and hc are the normal and critical depths, given, respectively, by
1=3
Q2
hN ¼ ;
So C 2 b2
2 1=3 ð3:57Þ
Q
hc ¼ :
gb2
du So 1 u3
¼ ; ð3:58Þ
dx hN 1 nu3
that is,
Zu
So 1
x ¼ u ð1 nÞ du ¼ u ð1 nÞUðuÞ; ð3:61Þ
hN 1 u3
0
3.5 Computation of Steady Flow Profiles 165
Zu
1
UðuÞ ¼ du: ð3:62Þ
1 u3
0
or in the alternative form (Bresse 1860, p. 249) quoted by Chow (1959, p. 258) and
Jaeger (1956, p. 72)
" #
1 u2 þ u þ 1 1 1 2u þ 1
UðuÞ ¼ ln cot þ C1 : ð3:65Þ
6 ðu 1Þ2 31=2 31=2
Applying Eq. (3.66) involves an inverse computation: starting with the known
point (xi, hi) an increment in the water depth is defined such that
hi+1 = hi ± Δh. With the known values of normalized water depths ui and ui+1,
U(u) is evaluated at both points, thereby applying Eq. (3.66) to obtain the coor-
dinate xi+1 where the depth is hi+1. This process is conducted as many times as
2
The readerhis warned i that Bresse’s
solution
appeared mistyped in many publications in the form
þuþ1
tan1 2u3 þ 1 þ C1 , which is obviously not the solution. Jan (2014,
2 1=2
UðuÞ ¼ 16 ln u ðu1Þ2 31=2
1
p. 24) detailed the integration process step-by-step. The typo appears to originate from
C. J. Posey in Rouse (1950, p. 613).
166 3 Computation of Steady Gradually-Varied Flows
desired to get the free surface profile. Note that although the computational
sequence looks like a numerical scheme, it is not. The solution is analytical and
therefore the value of Δh arbitrary, since Eq. (3.66) is only a relation between two
analytical points of the flow profile.
In this section, a numerical method based on the integral form of the energy
balance, namely Eq. (3.51), is presented. The technique is called standard step
method (Chow 1959) and consists in dividing the channel reach into a number of
segments with their boundaries located at selected positions, and then solve the
algebraic equation for the unknown flow depth at one of the extremes. This tech-
nique is widely used in river hydraulics, and it is implemented in program
HEC-RAS (Brunner 2016). Here, we follow the presentation by Fread and
Harbaugh (1971). Consider a channel reach divided for simplicity’s sake into
segments of equal length Dx (Fig. 3.20).
Using the trapezoidal rule to evaluate the integral of the friction slope, one gets
xZþ Dx
1
Sf dx ¼ Sf 1 þ Sf 2 Dx; ð3:67Þ
2
x
1
E1 ¼ E2 So Dx þ ðSf 1 þ Sf 2 ÞDx; ð3:68Þ
2
Q2 Q2 1
zb1 þ h1 þ 2
¼ zb2 þ h2 þ þ ðSf 1 þ Sf 2 ÞDx; ð3:69Þ
2gA1 2gA22 2
Q2 Dx n2 Q2
f ðh1 Þ h1 þ þ F ¼ 0; ð3:70Þ
2gA21 2 A2 R4=3
1 h1
Q2 Dx n2 Q2
F ¼ So Dx h2 : ð3:71Þ
2gA22 2 A2 R4=3
2 h2
The solution of Eq. (3.70) is conducted here using the Newton–Raphson method as
fk
hk1 þ 1 ¼ hk1 ; ð3:72Þ
ðdf =dh1 Þk
with k as the iteration index. Since Dx is usually small, the initial value of h1 is
taken as h2. For the computation of a supercritical flow profile (F > 1), the upstream
water depth is h2 and the unknown water depth h1 is located at the downstream
168 3 Computation of Steady Gradually-Varied Flows
boundary. The equations presented are thus applied just taking a negative value of
Dx to conduct computations in the flow direction.
The derivative needed in Eq. (3.72) is
!
df d Q2 Dx n2 Q2 d 1
¼ h1 þ 2
þF ¼ E1 DxSf 1 : ð3:73Þ
dh1 dh1 2gA1 2 A2 R4=3 dh1 2
1 h1
The derivative of E is
dE1 Q2
¼ 1 F21 ¼ 1 B1 3 ; ð3:74Þ
dh1 gA1
Inserting Eqs. (3.74) and (3.75) into Eq. (3.73), the final result is
df Q2 1 10B1 4 1 dp1
¼ 1 B1 3 DxSf 1 þ : ð3:76Þ
dh1 gA1 2 3A1 3 p1 dh1
1
E2 E1 ¼ So Dx Sf 1 þ Sf 2 Dx: ð3:77Þ
2
Starting with the known point (xi, hi), an increment in the water depth is defined
such that hi+1 = hi ± Δh. Assume that E = E(h) and Sf = Sf(h) only, i.e., the
channel is prismatic. Thus, one can write
E2 E1
Dx ¼ x2 x1 ¼ : ð3:78Þ
So 12 Sf 1 þ Sf 2
From this equation, we can compute the position x2 where the depth will be h2,
given that the right-hand side of Eq. (3.78) is only a function of h1 and h2. This
method is called direct step method and was originally devised by Bélanger (1828).
A limitation is that it can only be used for prismatic channels, and that it does not
permit to get the solution for the flow depth at selected locations, which is the usual
case in practice. Thus, aside from its academic interest, is not a practical method.
For the solution of the GVF-ODE [Eq. (3.49)], the computational domain is divided
into a number of reaches of length Dx (Fig. 3.21). If the spatial derivative of the
flow depth is discretized using a forward finite-difference
dh hi þ 1 hi
¼ ; ð3:79Þ
dx Dx
Thus, the flow solution is computed on the basis of the tangent to the known point
(xi, hi) (Fig. 3.22). This equation has the advantage of being extremely simple to
apply, but the truncation error is O(Dx2), and it can destroy the solution if Dx is
large. It is noted that we are including in the solution only the terms of power Dx of
a Taylor series, e.g.,
2 3
dh dh 1 2 d h 1
hi þ 1 ¼ hi þ Dx þ 2
ðDxÞ þ ðDxÞ3 þ : ð3:81Þ
dx i dx i 2 dx3 i 6
The scheme is therefore first-order accurate only. It applies in practice, but extreme
care is needed to avoid an unacceptable growth of the numerical error in the
solution.
The family of Runge–Kutta schemes considers the Taylor series Eq. (3.81), sub-
stituting the use of derivatives of order higher than one by more evaluations of the
first derivative within the interval. These schemes use an updating explicit formula
of type
hi þ 1 ¼ hi þ f Dx; ð3:82Þ
3.5 Computation of Steady Flow Profiles 171
where the average value of f is determined differently, depending on the order of the
scheme, and, thus, of the number of terms considered in the Taylor series.
A fourth-order scheme yields (Apelt 1971; Chaudhry 2008)
1
f ¼ ðk1 þ 2k2 þ 2k3 þ k4 Þ; ð3:83Þ
6
with
k1 ¼ f ðxi ; hi Þ;
1 1
k2 ¼ f xi þ Dx; hi þ k1 Dx ;
2 2
ð3:84Þ
1 1
k3 ¼ f xi þ Dx; hi þ k2 Dx ;
2 2
k4 ¼ f ðxi þ Dx; hi þ k3 DxÞ:
The computational process is depicted in Fig. 3.23, where the four evaluations of
the slope are seen. The truncation error of the scheme is O(Dx5) (Apelt 1971). This
scheme is suitable for open channel flows (Katopodes 2019). Apelt (1971) found
that this is an accurate and stable method to integrate Eq. (3.49) in the up- and
downstream directions regardless of whether the flow is subcritical or supercritical.
Equation (3.80) is simply the first-order Runge–Kutta method.
Another relevant method uses the integral form of the GVF-ODE, e.g.,
Zxi þ 1
hi þ 1 ¼ hi þ f ðx; hÞdx: ð3:85Þ
xi
The problem is to evaluate the integral numerically. Using the trapezoidal rule, one
gets the implicit equation in the depth hi+1 as (Apelt 1971; Chaudhry 2008)
1
hi þ 1 ¼ hi þ Dx½f ðxi ; hi Þ þ f ðxi þ 1 ; hi þ 1 Þ: ð3:86Þ
2
This value is used to initiate the iterative corrector cycle, described below.
1
hki þþ11 ¼ hi þ Dx f ðxi ; hi Þ þ f xi þ 1 ; hkiþ 1 : ð3:88Þ
2
Iteration of hi+1 in the corrector cycle is stopped once the solution accuracy is
within a prescribed tolerance for two successive values of the iterated flow depths,
that is,
kþ1
hi þ 1 hkiþ 1
106 : ð3:89Þ
hkiþ 1
3.6 Applications
Fig. 3.24 Evaluation of fourth-order Runge–Kutta scheme in transition from mild to steep slope
174 3 Computation of Steady Gradually-Varied Flows
Fig. 3.25 Comparison of various numerical schemes to solve the GVF-ODE using an M2 profile
still small. This is one advantage of the standard step method, namely the possibility
of using larger space steps given its implicit character.
The GVF equation for a rectangular channel of steep slope is [Eq. (1.151)]
2
2 2
dh So Cf F2 ð1 þ S2o Þ2 So Cf gh3 ð1 þ So Þ
q
¼ ¼ ; ð3:90Þ
1 F2 ð1 þ S2o Þ
2
dx 1 gh
q
3 ð1 þ So Þ
2
where Cf is the skin friction coefficient. Note that under the small slope assumption
1 + S2o 1, so that Eq. (3.90) simplifies to Eq. (3.47). A steep physical slope is
176 3 Computation of Steady Gradually-Varied Flows
pb h
¼ : ð3:91Þ
qg 1 þ S2o
The test data of Hasumi (1931) for slopes of So = 1 and 1.732 were used to
check predictions by Eqs. (3.90) and (3.91). Equation (3.90) was solved with the
fourth-order Runge–Kutta scheme, available from the file steepslope_Runge
Kutta4order.xls in Chap. 12. The slope transition is composed of a horizontal reach
followed by a circular-shaped transition profile of R = 0.1 m that finishes in the
steep slope reach of 45° or 60° inclination angles. The channel width is
b = 0.402 m and the discharge Q = 0.04 m3/s. A value of Cf = 0.001 was used in
the simulations.
The first test for So = 1 is presented in Fig. 3.28a, where the boundary condition
taken to integrate Eq. (3.90) was the experimental point at the start of the flat
slope reach (xo = 0.070711 m, ho = 0.075 m). The elevation at this point is
zo = –0.02929 m. The computed free surface profile is compared with the
3.6 Applications 177
experiments by Hasumi (1931), showing a fair agreement down the slope. Once the
free surface is determined, the bottom pressure head was computed from Eq. (3.91),
showing again good agreement with experiments down the slope. The second test
for So = 1.732 is presented in Fig. 3.28b, with the boundary condition set as the
experimental point at the start of the flat slope reach (xo = 0.086602 m, ho =
0.081 m, zo = –0.05 m). A comparison of numerical predictions and experimental
observations in Fig. 3.28b shows again a good agreement down the slope.
The importance of the mild slope assumption 1 + S2o 1 is tested in Fig. 3.29,
where the tests are repeated this time including the solution of Eq. (3.90) neglecting
the slope corrections. The effect of the correction on the free surface for the 45°
slope is moderate (Fig. 3.29a), but obviously, the bottom pressure is poorly pre-
dicted, given that for the mild slope case it is
pb
¼ h: ð3:92Þ
qg
178 3 Computation of Steady Gradually-Varied Flows
Bottom pressures are not considered in Fig. 3.29, therefore. The impact of the slope
correction on the free surface for the 60° slope is depicted in Fig. 3.29b, showing in
that case important deviations.
1=2
Sf 2=3
Ui ¼ Rhi : ð3:93Þ
ni
To compute the hydraulic radius of a subsection, only solid perimeters are con-
sidered. The free surface profile is determined from the integral energy balance as
(Jain 2001)
Q2 Q2 1
zb1 þ h1 þ a1 ¼ z b2 þ h 2 þ a 2 þ Sf 1 þ Sf 2 Dx; ð3:94Þ
2gA21 2gA22 2
where the friction slope Sf and Coriolis coefficient a shall be determined at each
section. The total discharge in the cross-section is (Chow 1959; Jain 2001)
X
N
1=2
X
N
1=2
Q¼ Ui Ai ¼ Sf Ki ¼ KSf ; ð3:95Þ
i¼1 i¼1
where
1 2=3
Ki ¼ Ai Rhi : ð3:96Þ
ni
Inserting Eqs. (3.95) and (3.97) in Eq. (3.94), it can be solved by the standard step
method (Jain 2001). The determination of control sections, especially critical depth
controls, is a complex task, given the multiple critical depths in the specific energy
diagram (Sturm 2001). Further, various theoretically possible free surface profiles
in compound channel flow are not yet experimentally verified.
References
Apelt, C. J. (1971). Numerical integration of the equation of gradually varied and spatially varied
flow. In: Proceedings of 4th Australasian Conference on Hydraulics and Fluid Mechanics,
Monash University, Melbourne, Australia (pp. 146–153).
Bakhmeteff, B. A. (1912). O нepaвнoмepнoм движeнии жидкocти в oткpытoм pycлe [Varied
flow of liquids in open channels]. St Petersburg, Russia (in Russian).
Bakhmeteff, B. A. (1932). Hydraulics of open channels. New York: McGraw-Hill.
Bélanger, J. B. (1828). Essai sur la solution numérique de quelques problèmes relatifs au
mouvement permanent des eaux courantes [On the numerical solution of some steady water
flow problems]. Carilian-Goeury, Paris (in French).
Bresse, J. (1860). Cours de mécanique appliquée, 2ème partie: Hydraulique [Lecture notes on
applied mechanics, Part 2, Hydraulics]. Paris: Mallet-Bachelier (in French).
Brunner, G. W. (2016). HEC-RAS River Analysis System Hydraulic Reference Manual. Report
CPD69, Version 5.0, US Army Corps of Engineers, Hydrologic Engineering Center (HEC),
Davis, CA.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Chapra, S. C., & Canale, R. P. (2010). Numerical methods for engineers (6th ed.). New York:
McGraw-Hill.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). New York: Springer.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Fenton, J. D. (2010). Calculating resistance to flow in open channels. Technical Report,
Alternative Hydraulics Paper 2, Vienna University of Technology, Vienna, Austria.
Fread, D. L., & Harbaugh, T. E. (1971). Open-channel profiles by Newton’s iteration technique.
Journal of Hydrology, 13, 78–80.
Hager, W. H. (2010). Wastewater hydraulics: Theory and practice (2nd ed.). Berlin: Springer.
Hasumi, M. (1931). Untersuchungen über die Verteilung der hydrostatischen Drücke an
Wehrkronen und -Rücken von Überfallwehren infolge des abstürzenden Wassers [Studies on
the distribution of hydrostatic pressure distributions at overflows due to water flow]. Journal of
the Department of Agriculture, Kyushu Imperial University 3(4), 1–97 (in German).
Henderson, F. M. (1966). Open channel flow. New York: MacMillan Co.
Hoffman, J. D. (2001). Numerical methods for engineers and scientists (2nd ed.). New York:
Marcel Dekker.
Jaeger, C. (1956). Engineering fluid mechanics. Edinburgh: Blackie and Son.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Jan, C.-D. (2014). Gradually-varied flow profiles in open channels: Analytical solutions by using
Gaussian hypergeometric function. Advances in Geophysical and Environmental Mechanics
and Mathematics. Berlin: Springer.
Jeppson, R. (2011). Open channel flow: Numerical methods and computer applications. Boca
Raton: CRC Press.
Katopodes, N. D. (2019). Free surface flow: Shallow-water dynamics. Oxford, UK:
Butterworth-Heinemann.
References 181
Lee, M. T., Babbitt, H. E., & Baumann, E. R. (1952). Gradually varied flow in uniform channels
on mild slopes. Engineering Experiment Station Bulletin Nº 404, University of Illinois at
Urbana Champaign, College of Engineering, Engineering Experiment Station.
Montes, J. S. (1998). Hydraulics of open channel flow. Reston, VA: ASCE.
Puertas, J., & Sánchez, M. (2001). Apuntes de Hidráulica de canales [Open channel hydraulics
lecture notes]. Civil Engineering School, University da Coruña, Spain (in Spanish).
Rouse, H. (1938). Fluid mechanics for hydraulic engineers. New York: McGraw-Hill.
Rouse, H. (1950). Engineering hydraulics. New York: Wiley.
Rouse, H. (1965). Critical analysis of open-channel resistance. Journal of the Hydraulics Division,
ASCE, 91(HY4), 1–25.
Sturm, T. W. (2001). Open channel hydraulics. New York: McGraw-Hill.
Subramanya, K. (1986). Flow in open channels. New Delhi: Tata McGraw-Hill.
White, F. M. (1991). Viscous fluid flow. New York: McGraw-Hill.
White, F. M. (2003). Fluid mechanics. New York: McGraw-Hill.
Woodward, S. M., & Posey, C. J. (1941). Hydraulics of steady flow in open channels. New York:
Wiley.
Yen, B. C. (1973). Open-channel flow equations revisited. Journal of the Engineering Mechanics
Division, ASCE, 99(EM5), 979–1009.
Yen, B. C. (1991). Hydraulic resistance in open channels. In B. C. Yen (Ed.), Channel flow
resistance: Centennial of Manning’s formula (pp. 1–135). Highlands Ranch, USA: Water
Resources Publications.
Yen, B. C. (2002). Open channel flow resistance. Journal of Hydraulic Engineering, 128(1),
20–39.
Chapter 4
Computation of Steady Transcritical
Open Channel Flows
4.1 Introduction
There are two possible transitional open channels flows, as described in Chap. 2:
the transition from subcritical ðF\1Þ to supercritical ðF [ 1Þ flows, and the
transition from supercritical ðF [ 1Þ to subcritical ðF\1Þ flows. Both of these flow
types are fundamentally different. The transition from subcritical to supercritical
flows is smooth and continuous, as, for example, in weir flow (Fig. 4.1a). The
transition from supercritical to subcritical flows, however, is abrupt, encompassing
a highly turbulent free surface profile (Fig. 4.1b). This chapter details the compu-
tation of transitional flow profiles. Computations are then illustrated and compared
with experimental observations.
Fig. 4.1 Transitions from a subcritical ðF\1Þ to supercritical ðF [ 1Þ flows over a weir,
b supercritical ðF [ 1Þ to subcritical ðF\1Þ flow in a hydraulic jump, with F as the Froude
number
Fig. 4.2 Transition from a mild to a steep bottom slope with a gradual bottom transition curve:
Formation of a critical point, where the flow depth equals the critical depth and the friction slope is
exactly equal to the bottom slope at this section
(1938), recently reviewed by Hager and Castro-Orgaz (2016). Between these two
reaches, the transitional bed profile is assumed to be smooth and continuous up to
second derivatives. The gradually varied flow equation is thus for a rectangular,
prismatic channel (Henderson 1966).
4.2 Transition from Sub- to Supercritical Flow 185
Pierre Massé was born on January 13, 1898, in Paris, passing away there
aged nearly 90 years old on December 15, 1987. He obtained in 1928 the
civil engineering degree from Ecole des Ponts et Chaussées, entering sub-
sequently in the electric industry of France, today’s Electricité de France
(EDF). He submitted in 1935 his Ph.D. thesis on variable regimes in fluvial
hydraulics, in which his mathematical knowledge was developed. Further
works are related to tidal flows and flows in mountainous rivers, in which
aspects of supercritical flow are fundamental, including the transition to
subcritical flow. His important work on the hydraulic jump and the free
surface profile in channels of variable bottom slope published in 1938 was
awarded the Prix Caméré from the French Académie des Sciences. He in
parallel was active in the hydropower industry, collaborating with the great
French dam designer André Coyne (1891–1960) on the Portillon Dam in the
French Pyrenees, or the Chastang Dam on Dordogne River in Southern
France. He was appointed in 1946 EDF Director of equipment, and in 1948
Vice-Director EDF, presiding from 1965 to 1969 the EDF Administrative
Council. He became in 1955 Commandeur of the French Légion d’Honneur
and was in 1977 elected Member of Académie des Sciences. The latter dis-
tinction is normally given to scientists only, shedding thus light on his great
career both in theoretical and applied engineering.
dzb
dh dx Sf So ð xÞ Sf
¼ ¼ ; ð4:1Þ
dx q2 1 F2
1 3
gh
where h is the flow depth, g the gravity acceleration, q the discharge per unit width,
F ¼ q=ðgh3 Þ1=2 the Froude number, Sf the friction slope, and So the bed slope,
defined as
186 4 Computation of Steady Transcritical Open Channel Flows
dzb
So ¼ : ð4:2Þ
dx
If there are not gates, weirs, or, in general, any other channel control along the
channel reach, the only possible singularity provoking a variation in the flow
conditions is the variable slope reach zb = zb(x). Far up- and downstream along the
uniform slopes, asymptotic conditions will be settled, with the flow depth
approaching the corresponding uniform flow depth hN. The flow gradually adjusts
from the upstream subcritical uniform depth (hN > hc) to the downstream super-
critical uniform depth (hN < hc). The free surface profile will thus cut somewhere
the critical depth line to change from sub- to supercritical conditions. The flow
conditions at the critical flow section are investigated following Massé (1938) and
Henderson (1966). Rewrite Eq. (4.1) as
dh
1 F2 ¼ S o S f : ð4:3Þ
dx
If F ¼ 1 is set into Eq. (4.3), then So = Sf. As the flow gradually changes from
upstream subcritical flow to downstream supercritical flow, dh=dx 6¼ 0. Thus, at the
section where the flow depth equals the critical flow depth, the friction slope equals
the local bottom slope, and, obviously, the water surface slope at this section is
different from zero (Henderson 1966; Hager and Castro-Orgaz 2016). The point
satisfying these conditions is referred to as singular point, acting as a control section
of the flow. Starting at the singular point, the subcritical branch of the free surface
profile is computed in the upstream direction, whereas the supercritical portion is
computed in the downstream direction. A particular case of notable interest is the
transition from an adverse to a steep slope; it is the case occurring in weir flow
(Fig. 4.3) (Puertas and Sánchez 2001).
Fig. 4.3 Formation of a critical point in ideal fluid flow along the transition from an adverse to a
steep positive slope (adapted from Puertas and Sánchez 2001)
4.2 Transition from Sub- to Supercritical Flow 187
The fluid flow is assumed to be ideal given the short transitional length, and,
thus, Sf = 0. Equation (4.3) then produces
dh
1 F2 ¼ S o : ð4:4Þ
dx
Focus on the weir crest, where So = 0. Inserting this in Eq. (4.4) yields
dh
1 F2 ¼ 0: ð4:5Þ
dx
This identity is verified for the obvious case dh/dx = 0. This case is described in
Chap. 2 and corresponds to either fully sub- or supercritical flows over the weir
profile. However, the identity is also verified with F ¼ 1, corresponding to critical
flow at the weir crest with dh/dx < 0, given that the water accelerates from sub- to
supercritical flows, so that the flow depth must decrease in the flow direction
(Fig. 4.3).
If F ¼ 1, then Eq. (4.4) produces the indetermination dh/dx = 0/0 at the weir crest.
This singularity in the equation of motion is removed by applying L’Hospital’s rule
to Eq. (4.4) (Massé 1938; Escoffier 1958; Hager and Castro-Orgaz 2016). This
technique to remove flow depth gradients of the kind 0/0 on the shallow water
steady-state equations is known as the singular point method in open channel
hydraulics (Chow 1959; Montes 1998; Hager 1985, 2010). It originates from the
work of Poincaré (1881) on ODE equations and was applied to open channel
transition flow problems by Massé (1938), Escoffier (1958), Iwasa (1958), Wilson
(1969), and Chen and Dracos (1996). However, this method is questioned in the
literature given that the argument still prevails that Eq. (4.4) is invalid for h = hc
given the existence of non-hydrostatic pressure distribution. The gradually varied
flow model is mathematically valid at the critical depth, but, physically, it is
inaccurate if the free surface curvature is high, given that then the hydrostatic
pressure approximation is inadequate (Montes 1998; Castro-Orgaz and Chanson
2016). The singular point method is rarely explained in open channel flow books,
with Chow (1959), Henderson (1966) and Montes (1998) as notable exceptions.
However, mathematical books often describe it for general application in engi-
neering (i.e., von Kármán and Biot 1940).
Consider the weir flow depicted in Fig. 4.3. The boundary condition to integrate
Eq. (4.4) is the critical depth at the weir crest, h(x = 0) = hc = (q2/g)1/3, with the
origin of the x-coordinate at the weir crest. To integrate the ODE in the up- and
downstream directions using the Runge–Kutta method, the value of the free surface
188 4 Computation of Steady Transcritical Open Channel Flows
slope dh/dx at the crest must be determined (Chen and Dracos 1996). However, this
value is unknown, given the mathematical indetermination existing there.
Equation (4.4) is a steady-state version of the SWE momentum equation. However,
in steady gradually varied flow, it is also obtained by differentiation of the total
energy head H = zb + h + q2/(2gh2) as
d q2
zb þ h þ ¼ 0; ð4:6Þ
dx 2gh2
where energy losses are neglected given the short transitional reach. Its first dif-
ferential is
dzb dh q2 dh
þ ¼ 0; ð4:7Þ
dx dx gh3 dx
Inserting into Eq. (4.8) the singular point conditions at the weir crest
q2 dzb
¼ 1; ¼ 0; ð4:9Þ
gh3c dx
Note from this finding that the physically correct result implies a minus sign in front
of the square root function, corresponding to accelerating flow at the weir crest.
Further, given that hc is positive, imaginary numbers originate as a solution for
dh/dx if dz2b/dx2 is positive, meaning that a concave free surface profile is not
compatible with weir flow. Only a convex bottom profile (dz2b/dx2 < 0) produces a
transcritical free surface profile passing from sub- to supercritical flows, therefore.
The mathematical problem is thus the solution of the following ODE with critical
flow as the boundary condition:
4.2 Transition from Sub- to Supercritical Flow 189
2 8
>
> dzb
6 >
>
6 >
> dx if x 6¼ 0;
6 >
dh < 1 q
2
6
6 ODE: ¼
6 dx >> gh3
6 >
> 2 1=2 ð4:12Þ
6 >
> h d zb
6 >
:
c
if x ¼ 0;
6 3 dx2
6 2 1=3
6
4 q
Boundary condition: h ð x ¼ 0 Þ ¼ hc ¼ :
g
The transition from super- to subcritical flows is rapidly varied and, thus, cannot be
handled using the gradually varied flow theory. The pressure distribution can be
considered hydrostatic, but the velocity distribution is markedly non-uniform,
Reynolds stresses are extremely high and air entrainment provokes a two-phase
flow rather than a clear water flow (Fig. 4.1b). The flow is definitely not gradually
varied within a hydraulic jump, so Eq. (4.1) cannot be integrated to predict the free
surface profile of the hydraulic jump. Resort to more advanced approaches is
necessary for this task (Castro-Orgaz and Hager 2009), which is beyond the scope
of this book. The hydraulic jump is treated here using control volume equations,
with the purpose of relating the values of depth and velocity at the boundary
sections, where the flow is gradually varied (Fig. 4.4a).
Although the equation describing a hydraulic jump is presented in Chap. 2,
some developments are repeated here for convenience to introduce the hydraulic
jump as an element of free surface profile computations. Consider the hydraulic
jump in a horizontal and rectangular channel depicted in Fig. 4.4a; the control
volume with boundary sections 1 and 2, where the flow is super- and subcritical,
respectively, is used to formulate integral balances. At these boundaries, pressure is
hydrostatic and velocity uniform. Reynolds’ transport theorem states the conser-
vation of mass and momentum as (Liggett 1994; Sturm 2001; Chaudhry 2008),
ZZZ ZZ
d
qdv þ qðV nÞdA ¼ 0; ð4:13Þ
dt
CV CS
X ZZZ ZZ
d
f¼ qVdv þ qVðV nÞdA: ð4:14Þ
dt
CV CS
Here, V is the velocity vector, f a fluid force component, n the unit vector normal
to the control area A, v the volume, and q the water density. CV and CS are the
190 4 Computation of Steady Transcritical Open Channel Flows
Fig. 4.4 Hydraulic jump in a horizontal channel a control volume for application of Reynolds’
transport theorem, b approximate representation of hydraulic jump as a local discontinuity
(Dx = 0) in the shallow water hydraulics theory
control volume and its surrounding control surface. The boundary limits of SC are
flow sections 1 and 2. For steady 1D flow, these reduce to
q ¼ U 1 h1 ¼ U 2 h2 ; ð4:15Þ
X
Fx ¼ qqðU1 U2 Þ; ð4:16Þ
1 2
qg h2 h21 ¼ qqðU1 U2 Þ: ð4:17Þ
2
h1 h 1=2 i q2
h2 ¼ h1 ¼ 1 þ 1 þ 8F21 ; F21 ¼ : ð4:18Þ
2 gh31
4.3 Transition from Super- to Subcritical Flows 191
This relation is known as Bélanger’s equation for the hydraulic jump (Bélanger
1849). The water depth h1 is called sequent depth of h1. From Eq. (4.18), it is easily
verified that a decrease of h1 implies an increase of h1 . Note also that the sequent or
conjugate depth of the critical depth is itself.
Within the context of the shallow water hydraulic theory, the effect of length of
the hydraulic jump is neglected (Dx = 0). The hydraulic jump is therefore treated as
a discontinuity determined by Eq. (4.18) (Fig. 4.4b). Physically, the difference
between the transition from sub- to supercritical and that from super- to subcritical
flows is easily appreciated (Fig. 4.1): The former is smooth and continuous,
192 4 Computation of Steady Transcritical Open Channel Flows
whereas the latter is abrupt and highly turbulent. Mathematically, this is also stated
by the shallow water hydraulics theory: Ideal weir flow results in a mathematical
indetermination producing a smooth solution, whereas the hydraulic jump is
mathematically a discontinuity accompanied with significant energy losses.
For a given discharge q and supercritical flow depth h1, Eq. (4.18) defines the
required value of h2, namely h1 , to produce a steady hydraulic jump. If given h1, for
example, the actual tailwater level h2 is less than h1 , the shock front (h2 − h1) will
move in the downstream direction. The physical reason is that the momentum plus
pressure force of the incoming supercritical flow is larger than that at the tailwater
section. Thus, pressure forces and momentum are not in equilibrium, and the
unbalanced residual provokes an unsteady motion removing the shock in the
downstream direction. Only if h2 h1 , the jump is steady. The momentum equa-
tion of the hydraulic jump is rewritten as
h21 q2 h2 q2
þ ¼ 2þ ; ð4:19Þ
2 gh1 2 gh2
h2 q2
S1 ¼ S2 ; S¼ þ : ð4:20Þ
2 gh
Fig. 4.5 Moving hydraulic jump, or surge with h1 < h2, displacing in: a the downstream
direction, b the upstream direction, c with no displacement, that is, under steady conditions
(adapted from Puertas and Sánchez 2001)
It is now explained how to use Eq. (4.18) to determine the position of a hydraulic
jump in a free surface profile using a practical example. Consider a sluice gate
upstream of a free overfall in a mild slope channel. Two flow profiles are possible,
each determined by a different channel control: subcritical flow profile computed in
the upstream direction assuming critical flow at the brink section; supercritical flow
profile computed in the downstream direction imposed by a gate opening below the
194 4 Computation of Steady Transcritical Open Channel Flows
Fig. 4.6 Sub- and supercritical free surface profiles beyond a sluice gate
critical depth (Fig. 4.6). It must be determined which of the two profiles is the
actual physically valid solution, depending on which control is active, or, if both
controls are active (the brink section and the gate), the super- and subcritical
profiles must be assembled through a hydraulic jump. The computational process
encompasses the following steps:
(i) Compute the supercritical M3 curve h = h(x) starting at the gate section, or at
the vena contracta assuming Cc = 0.61, if contraction effects are accounted
for. The distance from the gate to the vena contracta is small and therefore
neglected in gradually varied flow computations. The computation must be
stopped ahead of reaching the critical depth line hc, given that dh/dx ! ∞.
(ii) After computation of the supercritical M3 profile, for each supercritical depth
h of this curve, compute the corresponding sequent depth from Bélanger’s
equation
hð x Þ h 1=2 i q2
hð x Þ ¼ 1 þ 1 þ 8F2 ; F2 ¼ : ð4:22Þ
2 g½ hð x Þ 3
Fig. 4.7 Free surface profile beyond a sluice gate with the formation of a hydraulic jump
Fig. 4.8 Subcritical free surface profile beyond a sluice gate: the gate is drowned
Therefore, the gate is drowned and the real free surface profile is the M2 curve, with
subcritical flow along the channel. In this case, the only active channel control is the
brink section. This case may occur if the channel beyond the sluice gate is long and
the normal depth is relatively high.
If the conjugate depth curve M3 of the M3 profile is above the M2 curve along
the entire channel reach (Fig. 4.9), no hydraulic jump is formed; the momentum
function of the M3 profile is larger than that for the M2 profile for all the channel
sections. Therefore, the hydraulic jump is fully removed from the channel and the
real free surface profile is the supercritical M3 curve. In this case, the only active
channel control is the gate, given that the flow is forced to exit from the channel at
the brink section at a high speed, and critical flow conditions cannot be settled
there, deactivating this as control. This case may occur if the channel beyond the
gate is short and the M3 curve is not reaching the critical depth line.
196 4 Computation of Steady Transcritical Open Channel Flows
Fig. 4.9 Supercritical free surface profile beyond a sluice gate: the hydraulic jump is rejected
from the channel reach
The steady transcritical free surface profile over a parabolic weir of bed shape
zb = 0.3216 − 0.54525x2 (m) was computed using Eq. (4.12). The numerical
model is implemented in the file parabolicweir_transcritical.xls of Chap. 12.
Equation (4.12) was solved using the fourth-order Runge–Kutta method (see
Chap. 3), and the corresponding sub- and supercritical branches of the free surface
profile were computed in the up- and downstream directions, respectively.
Theoretical predictions are compared in Fig. 4.10a, b with experimental data of
Blau (1963) for two runs corresponding to q = 0.51 and 0.151 m2/s, respectively.
In the first test, slight deviations between experiments and simulations are evident
given the effect of the non-hydrostatic pressure distribution, not accounted for by
the gradually varied flow theory. In the second run, however, predictions are in
excellent agreement with observations. At the weir crest, Eq. (4.11) was used to
remove the singularity of the equations of motion. This equation is compared in
Fig. 4.11 with the experimental measurements of Wilkinson (1974), confirming that
the result is not only theoretically sound, but also in agreement with experimental
observations (Castro-Orgaz and Chanson 2016).
Figure 4.12 contains the experimentalh data of Sivakumaran
i et al. (1983) for a
Gaussian hump of profile zb ¼ 20 exp 0:5ðx=24Þ2 (cm) for two test cases.
Simulations were conducted using Eq. (4.12) as previously described for the
parabolic weir. This test is implemented in the file Gaussianhump_transcritical.xls
in Chap. 12. The computed solution is presented for both cases and compared in
Fig. 4.12a, b with the experiments. The departure between simulations and
4.4 Computational Examples 197
experiments for the test case of Fig. 4.12a (q = 0.111 m2/s) indicates that the effect
of the non-hydrostatic pressure is notable as the flow passes from sub- to super-
critical. For the test case of Fig. 4.12b (q = 0.0359 m2/s), deviations between
numerical results and experiments are small.
198 4 Computation of Steady Transcritical Open Channel Flows
Fig. 4.13 Comparison of numerical simulation with experimental results (Gharangik and
Chaudhry 1991) for flow profile with steady hydraulic jump
the sequent depth curve fixes the position of the hydraulic jump. The computational
results are displayed in Fig. 4.13. A comparison of the transcritical free surface profile
with experimental data by Gharangik and Chaudhry (1991) in Fig. 4.13 shows good
agreement. Note that the finite length of the jump produced by turbulence is a feature
not accounted for in the theoretical predictions.
References
Bélanger, J. B. (1849). Notes sur le Cours d’Hydraulique (Notes on the Course in Hydraulics).
Mém. Ecole Nat. Ponts et Chaussées, Paris, France, session 1849–1850, 222 p (in French).
Blau, E. (1963). Der Abfluss und die hydraulische Energieverteilung über einer parabelförmigen
Wehrschwelle (Distributions of discharge and energy over a parabolic-shaped weir].
Mitteilung, 7, 5–72 Forschungsanstalt für Schiffahrt, Wasser- und Grundbau, Berlin (in
German).
Castro-Orgaz, O., & Hager, W. H. (2009). Classical hydraulic jump: Basic flow features. Journal
of Hydraulic Research, 47(6), 744–754.
Castro-Orgaz, O., & Chanson, H. (2016). Minimum specific energy and transcritical flow in
unsteady open channel flow. Journal Irrigation and Drainage Engineering, 142(1), 04015030.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). New York: Springer.
Chen, J., & Dracos, T. (1996). Water surface slope at critical controls in open channel flow.
Journal of Hydraulic Research, 34(4), 517–536.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Escoffier, F. F. (1958). Transition profiles in nonuniform channels. Transactions ASCE, 123, 43–
56.
Gharangik, A. M., & Chaudhry, M. H. (1991). Numerical simulation of hydraulic jump. Journal of
Hydraulic Engineering, 117(9), 1195–1211.
Hager, W. H. (1985). Critical flow condition in open channel hydraulics. Acta Mechanica, 54(3–
4), 157–179.
Hager, W. H. (2010). Wastewater hydraulics: Theory and practice (2nd ed.). Berlin: Springer.
Hager, W. H., & Castro-Orgaz, O. (2016). Critical flow in open channel hydraulics: From Boess to
De Marchi. Journal of Hydraulic Engineering, 142(1), 02515003.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan.
200 4 Computation of Steady Transcritical Open Channel Flows
5.1 Introduction
Flow conditions in rivers are usually unsteady (Cunge et al. 1980; Sturm 2001;
Chaudhry 2008), even though steady flow computations are conducted in engi-
neering applications, as for determining an inundation area in hydrological studies.
Further, the operation of man-made canals to control the water released from
reservoirs implies unsteady maneuvers resulting in transient open channel flows
(Chaudhry 2008). Therefore, real flow computations in canals and rivers require the
solution of the Saint-Venant equations or the shallow water equations (SWE). In
this chapter, these are presented and their basic continuous and discontinuous
unsteady flow solutions are discussed.
@h @ ðUhÞ @ ðVhÞ
þ þ ¼ 0;
@t @x @y
@ ðUhÞ @ 2 @ @zs 1
þ U h þ ðUVhÞ ¼ gh ðsxz Þb ; ð5:1Þ
@t @x @y @x q
@ ðVhÞ @ 2 @ @zs 1
þ V h þ ðVUhÞ ¼ gh syz b :
@t @y @x @y q
Here, (U, V) are the depth-averaged velocity components in the (x, y) horizontal
directions, h is the flow depth, zs the free surface elevation, g the gravity acceler-
ation, q the water density, and s denotes here tangential stresses. Most river flows
can be considered one-dimensional (1D), given that the x-direction is predominant
(Fig. 5.1). Therefore, the transverse water surface slope can be assumed horizontal.
To obtain 1D flow equations for an arbitrary cross section, the vertically averaged
continuity and x-momentum equations are laterally integrated in the y-direction (see
Chap. 1). The 2D depth-averaged model is then reduced to a 1D, sectional inte-
grated formulation. The river is therefore basically treated as a 1D stream tube.
Using Leibniz’s rule and imposing that the free surface slope is independent of the
y-direction (horizontal water surface), the lateral integration of Eqs. (5.1)1 and
(5.1)2 produces the 1D Saint-Venant equations (De Saint-Venant 1871) for an
arbitrary cross section (see Chap. 1)
@A @Q
þ ¼ 0;
@t @x
2 ð5:2Þ
@Q @ Q @zs so
þ ¼ gA gA :
@t @x A @x qgRh
Here, Q is the discharge, A the flow area, Rh = A/p the hydraulic radius, p the
wetted perimeter, and so the average boundary shear stress acting on the wetted
perimeter. The free surface elevation zs(x, t) corresponds at any vertical line of the
cross section to the sum of the bed elevation and the local water depth. The standard
approach is to use the thalweg (lowest point) to define the flow depth h and bed
elevation of the longitudinal river profile (Cunge et al. 1980) (Fig. 5.1). If the
channel is prismatic, Eq. (5.2) can be recast as (Sturm 2001; Jain 2001; Chaudhry
2008)
@A @Q
þ ¼ 0;
@t @x ð5:3Þ
@Q @ @zb
þ QU þ gAh ¼ gA Sf ;
@t @x @x
where U = Q/A, h is the flow depth along the thalweg, Sf ¼ so =ðqgRh Þ the friction
slope, bars indicate the depth of centroid below the free surface, and @zb =@x is the
longitudinal slope taking as reference the thalweg.
The x-coordinate is usually curvilinear, following the thalweg path in the hori-
zontal plane (x, y). Therefore, a river must not have severe curvatures in plan to
apply Eq. (5.3). Otherwise, curvature terms must be introduced to transform the
Cartesian system to curvilinear coordinates following the river plan, e.g., at river
meanders. For the basic case of a rectangular cross section, and with q = Q/b as the
unit discharge and b as the channel width, Eq. (5.3) reduces to
@h @q
þ ¼ 0;
@t @x
2 ð5:4Þ
@q @ q h2 @zb
þ þg ¼ gh Sf :
@t @x h 2 @x
Equation (5.4) is widely used in this book to introduce numerical methods. Note
that the differential form of the SWE only applies to zones of the (x, t) plane where
the dependent variables U(x, t) and h(x, t) are single-valued and smooth (Cunge
et al. 1980).
5.3.1 Introduction
A special feature of the Saint-Venant equations is that they admit both continuous
or gradually varied flow solutions, as well as discontinuous solutions (rapidly
varied flow zones), involving abrupt changes in the flow variables, even if initial
and boundary conditions are continuous (Cunge et al. 1980). Using shock-capturing
methods (Toro 2001), it is possible to simulate with accuracy both continuous and
discontinuous waves, where flow variables are not single-valued. To catch these
discontinuous solutions, the integral form of the SWE must be used (Cunge et al.
1980; Toro 2001). This is equivalent to apply the Reynolds transport theorem
(Chaudhry 2008, 2014; White 2009) to the fluid properties mass and momentum in
a finite control volume.
204 5 Unsteady Open Channel Flows: Basic Solutions
An important discontinuous unsteady open channel flow is the surge (Fig. 5.2),
corresponding to a traveling wave separating abruptly two open channel flow
portions of different depth and velocity. In this section, positive surges are
The celerity of wave propagation in gradually varied flow zones, where solutions
are smooth and continuous, is c = (gh)1/2 (see Chap. 2). This is the celerity of an
elementary gravity wave in a hydrostatic field. However, in a surge, the wave height
is finite, resulting in a different propagation celerity. As will be demonstrated, a
surge or bore can even propagate into the upstream direction in a supercritical flow,
something impossible within the framework of small gravity waves. Here, the basic
equations governing the movement of a surge will be developed considering
1
In computational fluid dynamics (CFD), the analogous discontinuous wave stemming from the
Euler equations for gas dynamics is called shock wave (Toro 2001, 2009; LeVeque 2002),
although there are also contact or shear waves. The hydraulic shock wave is simply referred to as
positive surge (Jain 2001; Chanson 2004; Chaudhry 2008).
206 5 Unsteady Open Channel Flows: Basic Solutions
Fig. 5.4 A tidal bore on Qiantang River at Yanguan, China, on 7/Sept/2013. Note the breaking
front propagating upstream (Photo courtesy of Prof. H. Chanson)
relations defining the surge height and velocity. The movement of the surge is
considered with respect to a fixed control volume in space and time, where
Reynolds’ transport theorem (Chow et al. 1988; Chaudhry 2008) is applied. It states
the conservation of any system property within a finite control volume. For mass
and momentum, these statements are, respectively (Liggett 1994; Chaudhry 2008),
ZZZ ZZ
d
qdv þ qðV nÞdA ¼ 0; ð5:5Þ
dt
CV CS
X ZZZ ZZ
d
f¼ qVdv þ qVðV nÞdA: ð5:6Þ
dt
CV CS
Here, V is the velocity vector, f a fluid force component, n the unit vector normal to
the control area A, v the volume, and q the water density. CV and CS are the control
volume and its surrounding control surface. The boundary limits of SC are flow
sections 1 and 2. The rate of temporal variation of water mass stored in the control
volume is after time Dt
ZZZ ZZZ
d d dv
qdv ¼ q dv ¼ q ¼ qVw ðA2 A1 Þ; ð5:7Þ
dt dt dt
CV CV
where A1 and A2 are the flow areas at sections 1 and 2. The net mass flow across the
contour of the control volume is
ZZ ZZ
qðV nÞdA ¼ qUdA ¼ qU1 A1 qU2 A2 : ð5:8Þ
CS CS
or
On inspecting Eq. (5.10), it is noted that the continuity equation across a surge is
simply obtained by superimposing a constant velocity (−Vw) to the unsteady flow
velocities. The unsteady translation of the surge is thus reduced to a steady problem
in moving axes at the absolute surge velocity. Likewise, the fluid force balance
projected in the x-direction of surge propagation in the control volume of Fig. 5.5
reads
208 5 Unsteady Open Channel Flows: Basic Solutions
X
fx ¼ qg h2 A2 h1 A1 : ð5:11Þ
This identity demonstrates that the momentum balance across a surge can also be
reduced to a steady flow balance by superimposing the constant velocity (−Vw) to
the unsteady flow velocities. From Eq. (5.10), one gets
A1
ðU2 Vw Þ ¼ ðU1 Vw Þ: ð5:16Þ
A2
A1
g h2 A2 h1 A1 ¼ A1 U1 ðU1 Vw Þ A2 U2 ðU1 Vw Þ
A2
ð5:16Þ
¼ A1 U1 ðU1 Vw Þ A1 U2 ðU1 Vw Þ
¼ A1 ðU1 Vw ÞðU1 U2 Þ:
5.3 Discontinuous Solutions: Basic Equations of a Positive Surge 209
U1 A1 U2 A2
Vw ¼ ; ð5:17Þ
A1 A2
This result shows that a finite discontinuity travels with a celerity different from that
of small gravity waves
c ¼ ðghÞ1=2 : ð5:23Þ
210 5 Unsteady Open Channel Flows: Basic Solutions
An important feature of surges is that they can run up in a supercritical current. For
a surge propagating in the upstream direction (Fig. 5.6), the sign of Vw by appli-
cation of the fundamental Eq. (5.22) must be negative.
Consider Eq. (5.22) rewritten as
1=2
gh2
Vw ¼ U1 ð h1 þ h2 Þ : ð5:24Þ
2h1
To obtain values of Vw < 0, the “−” sign must be selected. Let us define the surge
jump as Dh = h2 − h1 and the dimensionless jump height as e = Dh/h1. Using these
definitions, Eq. (5.24) yields
1=2 1=2
h1 þ Dh 1þe
Vw ¼ U1 g ðh1 þ h1 þ DhÞ ¼ U1 gh1 ð2 þ eÞ ; ð5:25Þ
2h1 2
or
h e i1=2
Vw ¼ U1 ðgh1 Þ1=2 ð1 þ eÞ 1 þ : ð5:26Þ
2
This implies with F1 ¼ U1 =ðgh1 Þ1=2 as the approach flow Froude number
1=2
Vw 3 e2
¼ F1 1 þ e þ : ð5:27Þ
ðgh1 Þ1=2 2 2
Equation (5.27) shows that in a supercritical current with F1 [ 1, the surge will
move upstream (Vw < 0) if the surge jump height e is large enough. The minimum
height emin for a given F1 above which the surge begins to move in the upstream
direction is from Eq. (5.27), by setting Vw = 0, determined by
3 e2
F21 ¼ 1 þ emin þ min : ð5:28Þ
2 2
Vw
¼ F1 1: ð5:29Þ
ðgh1 Þ1=2
This is always positive for F1 [ 1, implying that these infinitesimal surges cannot
travel in the upstream direction in supercritical flows. For a small amplitude e 1,
Eq. (5.27) reads
Vw 3 1=2
¼ F1 1 þ e : ð5:30Þ
ðgh1 Þ1=2 2
This equation was presented by De Saint-Venant (1870). Using it, surges propa-
gating in the upstream are generated. However, the resulting values of e are not of
small amplitude and thus nonlinear terms. In this case, better revert to the general
Eq. (5.27). The limiting case Vw = 0 corresponds to a steady surge or hydraulic
jump (Fig. 5.7).
212 5 Unsteady Open Channel Flows: Basic Solutions
Fig. 5.7 Hydraulic jump or steady surge (photograph VAW, ETH Zurich)
Solving Eq. (5.31) for the sequent depth ratio h2/h1 produces Bélanger’s equation
for the hydraulic jump presented in Chap. 4 as [Eq. (4.18)]
h2 1 h 1=2 i
¼ 1 þ 8F21 1 : ð5:32Þ
h1 2
The general equation for a shock wave [Eq. (5.22)] moving to the right direction is
1=2
gh2
Vw ¼ U1 þ ð h1 þ h2 Þ : ð5:33Þ
2h1
g ð h1 h2 Þ
ðU2 U1 Þ2 ¼ ðh1 þ h2 Þ; ð5:34Þ
2 h 1 h2
which coupled with Eq. (5.33) yields Vw as function of the tailwater velocity U2
1=2
gh1
Vw ¼ U2 þ ð h1 þ h2 Þ : ð5:35Þ
2h2
214 5 Unsteady Open Channel Flows: Basic Solutions
Using the definition c2 = (gh2)1/2 one gets from this last expression
1=2
h1
Vw ¼ U2 þ ðgh2 Þ1=2 ð h1 þ h2 Þ ¼ U2 þ Kc2 : ð5:36Þ
2h22
Equation (5.36) is a “signal” celerity used to propagate shocks in the finite volume
method, to be described in Chap. 9.
@U @F
þ ¼ S: ð5:38Þ
@t @x
Here, U is the vector of unknowns, F the flux vector, and S the source term vector,
given with Cf as a friction coefficient by
h Uh 0
U¼ ; F¼ ; S¼ ð5:39Þ
Uh U 2 h þ 12 gh2 gh @z
@x
b
Cf U jU j
This is called conservative form of the SWE (Toro 2001; Chaudhry 2008). This
system of PDEs is only valid for the computation of continuous solutions, where
h and q are smooth and single-valued in the (x, t) plane. The solution U(x, t) of
Eq. (5.38) must be differentiable with respect to x and t. To compute discontinuous
flows, the PDEs must be integrated over a control volume to obtain its integral form
(Cunge 1975). This fact was already used in Sect. 5.3.2 to obtain the basic relations
of a surge. Here, the integral form is reproduced by integrating the conservative
form of the SWE over a control volume. For simplicity, source terms are dropped,
resulting in the inviscid version of the SWE over horizontal terrain, that is
5.4 Methods of Solution 215
@U @F
þ ¼ 0: ð5:40Þ
@t @x
This equation is now integrated over an arbitrary control volume CV, resulting in
ZZZ
@U @F
þ dv ¼ 0; ð5:41Þ
@t @x
CV
or,
ZZZ
@U
þ r F dv ¼ 0: ð5:42Þ
@t
CV
Applying Gauss’s divergence theorem to the flux term in Eq. (5.42) results in
ZZZ ZZ
@U
dv þ ðF nÞdA ¼ 0: ð5:43Þ
@t
CV CS
2
In applied mathematics, the solutions of the integral form of a system of conservation laws are
called weak solutions of the differential form, even though they may or may not be the traditional
solutions of it (Cunge 1975; Macdonald 1995). This is simply a notation criterion to avoid having
to refer to the integral form, but, clearly, the integral form is linked to the concept of weak solution.
If a weak solution is continuous, then it is a solution of the differential form and it is called genuine
solution.
216 5 Unsteady Open Channel Flows: Basic Solutions
@h @U @h
þh þU ¼ 0;
@t @x @x ð5:44Þ
@U @U @h
þU þg ¼ 0:
@t @x @x
This form of the SWE generally applies only to compute continuous solutions.
A shock-capturing method based on Eq. (5.43) can resolve without any special
treatment surges and continuous waves, but, if applied to Eq. (5.44), will produce
shocks of erroneous celerity (Cunge 1975; Toro 2001) (see Sect. 5.8). An alter-
native is to use the so-called shock tracking methods: in these techniques,
Eq. (5.44) are used to solve the continuous portions of the solution (usually
resorting to the method of characteristics, see below), and the shocks are tracked
and characterized with the surge conditions as internal boundary conditions (Lai
1986; Cunge et al. 1980; Montes 1998). These techniques are not in use, given that
in the presence of multiple shocks computations are extremely tedious. Thus, shock
tracking methods are not further considered here.
@h @U @h
þD þU ¼ 0;
@t @x @x
ð5:45Þ
1 @U U @U @h
þ þ ¼ So Sf :
g @t g @x @x
Here, So ¼ @zb =@x is the bottom slope, Sf the friction slope and D = A/B the
hydraulic depth, with B as free surface width. Summing of the two equations, after
multiplying the continuity equation by an undetermined variable k yields, after
re-arrangement,
@U @U @h g @h
þ ðU þ kDÞ þk þ Uþ ¼ g So Sf : ð5:46Þ
@t @x @t k @x
The total differentials of the functions U = U(x, t) and h = h(x, t) are defined as
DU @U @U dx
¼ þ ;
Dt @t @x dt ð5:47Þ
Dh @h @h dx
¼ þ :
Dt @t @x dt
Comparing with Eq. (5.46), the terms inside the brackets are exact total differentials
only if
218 5 Unsteady Open Channel Flows: Basic Solutions
dx g
¼ U þ kD ¼ U þ : ð5:48Þ
dt k
dx
¼ U c; ð5:50Þ
dt
c ¼ ðgDÞ1=2 : ð5:51Þ
D
ðU þ xÞ ¼ g So Sf ; ð5:52Þ
Dt
provided that
dx
¼ U þ c; ð5:53Þ
dt
and
D
ðU xÞ ¼ g So Sf ; ð5:54Þ
Dt
if
dx
¼ U c: ð5:55Þ
dt
Zh
x¼ kdh: ð5:56Þ
0
5.5 Method of Characteristics 219
For a rectangular cross section, the reduced equations are after elementary
manipulations
D dx
ðU 2cÞ ¼ g So Sf along ¼ U c;
Dt dt ð5:57Þ
D dx
ðU þ 2cÞ ¼ g So Sf along ¼ U þ c:
Dt dt
Mathematically, this system of equations states that the total material derivatives of
(U ± 2c) are equal to g(So − Sf), but only along the paths in the x-t plane described
by the integrals of the ODEs dx/dt = U ± c. Therefore, the original SWE, a system
of PDEs valid at any point of the x-t plane, has been transformed into a system of
ODEs valid only along the paths given by the ODEs dx/dt = U ± c. Note that space
derivatives are no more in the equations, given that the x-coordinate is linked to
time by the integrals of dx/dt = U ± c.
Consider the case for which the source term is zero, that is, So = Sf = 0, resulting in
D dx
ðU 2cÞ ¼ 0 along ¼ U c;
Dt dt ð5:58Þ
D dx
ðU þ 2cÞ ¼ 0 along ¼ U þ c:
Dt dt
For inviscid flow over horizontal topography, the so-called Riemann invariants
(U ± 2c) are conserved along the paths in the x-t plane obtained upon integration of
the ODEs dx/dt = U ± c. These ODEs in the x-t plane are known as characteristic
curves. The family of curves given by the ODE dx/dt = U + c are known as for-
ward characteristics, while those given by dx/dt = U − c are known as backward
characteristics. In subcritical flow U < c (F\1), the forward and backward char-
acteristics curves in the x-t plane have positive and negative slopes, respectively
(Fig. 5.8a). In supercritical flow U > c (F [ 1), both the forward and backward
characteristics curves in the x-t plane have positive slopes (Fig. 5.8b). For critical
flow U = c (F ¼ 1), the backward characteristic in the x-t plane is a vertical line
(Fig. 5.8c).
Along the points on a forward characteristic,
D dx
ðU þ 2cÞ ¼ 0 ) U þ 2c ¼ const. along ¼ U þ c; ð5:59Þ
Dt dt
D dx
ðU 2cÞ ¼ 0 ) U 2c ¼ const. along ¼ U c: ð5:60Þ
Dt dt
220 5 Unsteady Open Channel Flows: Basic Solutions
Physically, this means that a perturbation generated at any point x at a given instant
of time t will propagate along the forward and backward characteristics curves
conserving along these paths the corresponding Riemann invariant. The propaga-
tion of any perturbation is therefore a transmission of information along the char-
acteristics: The non-equilibrium values h and U generated at the origin of the
perturbation will generate new non-equilibrium values at other sections at different
instants of time. In general, h and U are variable along the characteristics, which are
thus curved lines.
5.5 Method of Characteristics 221
channel. Typically, the discharge is constant and the variation of water depth in
space is computed by solving the steady-state version of the momentum equation
(see Chap. 3)
dq
¼ 0 ) q ¼ qo ¼ const.,
dx
dh So Sf So Sf ð5:63Þ
¼ 2 ¼ :
dx 1 Uc2 1 F2
Here, F is the Froude number and qo the inflow discharge. Further, two boundary
conditions are needed. These are the evolution of depth or velocity with time at
boundary sections of the channel reach. The sections to apply the boundary con-
ditions are different depending on whether the flow is sub- or supercritical. Consider
first the channel reach A–B (Fig. 5.11) with subcritical flow conditions throughout.
At the initial instant of time tn, flow conditions at points An, 2, 3, 4, 5, and Bn are
known. Inside the reach A–B, the forward and backward characteristics emerging
from points 3 and 4, respectively, intersect at point P. Conditions there are
Fig. 5.13 Example of initial and boundary condition requirements for subcritical flow
points 5 and 6. Thus, we do not need to prescribe any boundary condition at that
section. In supercritical flow, two boundary conditions are required at the upstream
section of the reach, and the flow is said to be controlled from upstream.
For illustrative purposes of initial and boundary data requirements, consider an
initially steady subcritical flow profile of type M1 shown in Fig. 5.13, formed at
the transition from a canal to a reservoir at its downstream end. Note that normal
depth is larger than the critical depth in this flow profile, as explained in Chap. 3.
The boundary sections of the canal reach are points 1 and 4. The initial condition
needed to solve the SWE is given by the M1 backwater profile, computed by
integration of Eq. (5.63). The typical boundary condition upstream of a subcritical
stream is the inflow discharge hydrograph, e.g., the function q = q(t) at point 1.
Likewise, a typical boundary condition at the downstream section of a subcritical
stream (point 4) is the water depth h in the form of rating curve q = q(h). This
rating curve may be determined by critical flow conditions, uniform flow con-
ditions, or a hydraulic element, e.g., a gate or a weir (see Chap. 2 for details of
governing equations for each boundary condition), depending on the physical
characteristics of the boundary. In our case, the water depth function in the
reservoir is assumed to be known. This relation is inverted to produce h = h(q),
and, with computed values of q(t) at point 4, the function h4 = h4(t) is described.
With these initial and boundary conditions (discharge upstream, water depth
5.5 Method of Characteristics 225
downstream), the SWE are numerically solved along the characteristics until a
new steady-state condition originates if the boundary conditions permit this
equilibrium solution.3
Thus, a subcritical flow is controlled from both up- and downstream, given that
perturbations at boundary sections propagate both up-and downstream of the reach.
The conditions at both boundary sections thus affect the flow inside of the compu-
tational reach. A typical boundary condition at the upstream section is the discharge,
e.g., a flood wave entering gradually into the reach or a sudden discharge increase
from one steady discharge to a new steady discharge. A typical boundary condition
at the downstream section is the water depth, controlled, for example, with
man-made structures, e.g., a gate or a weir. During unsteady flow, both boundary
conditions may depend on time, as explained. If a steady flow solution is physically
permitted by the boundary conditions, it is the asymptotic state produced by
unsteady computations based on the SWE (Macdonald 1995). For a steady sub-
critical flow, the upstream boundary condition q = q(t) reduces to qo = const. in the
entire channel reach (given that the continuity equation is dq/dx = 0), whereas the
downstream boundary condition h = h(t) transforms to h = const. at that section.
The steady form of the SWE is a first-order ODE, so only one boundary con-
dition is needed. This condition is, therefore, the downstream tailwater depth. The
upstream boundary condition of the SWE equations reduces simply to constant
discharge; i.e., the upstream constant discharge is conserved along the entire
channel reach. Based on this physical reasoning originating from the unsteady
SWE, it is possible to find a justification for the traditional rule stating “subcritical
steady flows are controlled from downstream”: Subcritical unsteady flows are
controlled from both up- and downstream, but in steady state, the upstream control
is “switched off.” Mathematically, the ODE for steady flow is a first-order equation
to be solved once a boundary condition is prescribed. From a mathematical
standpoint, this boundary condition is set either at the up- or downstream boundary
sections of the reach. Physically, a more logic choice in agreement with unsteady
SWE computations is to take the downstream boundary section. The same steady
computational rule states that “supercritical steady flows are controlled from
upstream”: Supercritical unsteady flows are controlled fully from upstream, but in
steady state, one of the upstream controls is “switched off.” At the upstream section,
both discharge and depth must be prescribed in the SWE. Following the former
discussion, in steady state, the discharge is a constant, so this is the boundary
condition that is automatically eliminated from the mathematical problem. The
water depth at the upstream boundary section of the reach remains as boundary
condition, in agreement with the steady-state computational rule.
3
Steady flow solutions can be generated using Eq. (5.63). To be of practical relevance, a steady
flow solution must be stable in time; that is, if the steady state is slightly perturbed, the flow
conditions must tend back again to the original steady state (Macdonald 1995). An example of
unstable steady flow solutions is presented when setting uniform flow on steep chutes, where roll
waves are generated (Stoker 1957).
226 5 Unsteady Open Channel Flows: Basic Solutions
The mathematical tools are now available to better understand the effect of a
stone (obstacle) inserted in an initially uniform and steady stream presented in
Chap. 2, given that it is in reality an unsteady gradually varied flow process. The
obstacle inserted into the initially uniform subcritical stream produces a gradual
variation of water levels (Fig. 5.14) (rapidly varied flow effects are overlooked for
the sake of simplicity). The upstream boundary condition at section A is the con-
stant discharge qo flowing initially in the uniform stream. The flow depths and
velocities within the reach evolve in time, but the discharge in the reach at section A
remains unaltered.
The downstream boundary condition at the vicinity of the stone (point B) is zero
discharge as long as the water depth is below the top of the stone. As the water
begins to pass over the stone, it acts like a round-crested weir, with critical depth
somewhere on the stone surface, close to the top of its round-crested surface
(Fig. 5.15).
Let q be the discharge at section B at any instant of time, the discharge equation
of the flow over the stone is with E as the specific energy over the crest
5.5 Method of Characteristics 227
3=2
2 3 1=2
q¼ gE : ð5:64Þ
3
q2
E ¼ Dþhþ : ð5:65Þ
2gh2
Thus, the boundary condition at point B is the function h = h(q) if h > D, obtained
by numerically inverting Eq. (5.66) using, e.g., the Newton–Raphson method. For
h < D, the boundary condition is simply q = 0.
With this reasoning, boundary conditions at points A and B are determined. The
net of characteristics is as shown in Fig. 5.14. Consider the backward characteristic
4-P. It transmits back in space, and forward in time, information on the perturbation
generated at point B. This information is transmitted back and back in space until at
a given instant of time the upstream water levels at A start to increase, in response
to this “information feedback.” After some time, a steady backwater profile is set
within the reach. Note that the discharge at section A was a constant during the
unsteady flow and that all the perturbation process was controlled by water levels at
B. Discharge at B is zero until the water begins to flow over the stone. Later, it
gradually increases until reaching the steady-state value qo that is the boundary
condition at the upstream section A. Therefore, the analogy with steady backwater
computations is evident: The discharge in the system is qo, and we set as boundary
condition at the tailwater section the depth h satisfying Eq. (5.66) for steady state,
that is,
For the supercritical case (Fig. 5.16), the upstream boundary condition is the
constant discharge qo flowing initially in the uniform stream. The other upstream
boundary condition needed is the uniform flow depth of the initial stream. There is
not any mathematical way of transmitting to the inlet (back in space) the pertur-
bation generated by the stone at the tailwater section, as demonstrated by the
characteristics plotted in Fig. 5.16. The flow profile within the channel reach is
uniform and steady, until it impacts the stone, jumping above it.
228 5 Unsteady Open Channel Flows: Basic Solutions
Fig. 5.17 Physical meaning of dx/dt in method of characteristics for a perturbation propagating
over initially steady non-uniform flow
the fact that the initial steady flow was non-uniform; in this case, the magnitude
dx/dt is not giving the velocity of propagation of a “visual” wave. It is, however,
the mathematical velocity at which the information of the perturbations in depth
and velocity travels along the characteristic lines. In specific cases, dx/dt can yield
a “visual” wave. It was demonstrated that a surge of infinitesimal height propa-
gates with celerity (gh)1/2 relative to the water flow. Thus, a surge of small height
will be visually observable traveling at a velocity close to this value. In this case,
in fact, both h and U remain constant along the characteristic lines. For a negative
surge propagating over an initially uniform flow, to be described below, it will be
shown that dx/dt also yields the wave celerity which can be visually observed.
230 5 Unsteady Open Channel Flows: Basic Solutions
dx x
¼ U o co : ð5:68Þ
dt t
This characteristic is a straight line given that both depth and velocity in the
undisturbed flow zone are constants. It can be demonstrated that if a member of a
C-family is a straight line, then all members of this family are also straight lines
(Liggett 1994; Jain 2001). Thus, for all C− characteristics, it can be stated that
dx
¼ U c ¼ const. ð5:69Þ
dt
U 2c ¼ const. ð5:70Þ
Thus, coupling Eqs. (5.69) and (5.70), it is seen that along the straight backward
characteristics both velocity and celerity are constants, e.g., (Fig. 5.18)
Thus, one can write, taking as reference the point at the intersection of a C− line
with the t-axis (e.g., point B in Fig. 5.18), where time is s,
UD ¼ Uo ;
ð5:75Þ
cD ¼ co :
UP ¼ UB ;
ð5:76Þ
cP ¼ cB ;
232 5 Unsteady Open Channel Flows: Basic Solutions
resulting from Eq. (5.74) after consideration of Eqs. (5.75) and (5.76) in
This relation states that the Riemann invariant U + 2c is a constant in the entire (x,
t) plane for the simple wave, that is,
dx
¼ U ð0; sÞ cð0; sÞ; ð5:79Þ
dt
dx
¼ Uo þ 2co 3cð0; sÞ: ð5:80Þ
dt
dx x
¼ : ð5:81Þ
dt t s
Consider a particular case of the simple wave problem. We seek a smooth wave
connecting two constant flow states A and B (Fig. 5.19). The simple wave solution
will be applied assuming the particular case where all C− characteristics are cen-
tered at the origin O, resulting in s = 0. The two edges of the wave will move at
velocities given by the corresponding backward characteristics. This wave is
referred to as rarefaction wave (Toro 2001), and it is a centered simple wave
solution of the SWE (Stoker 1957; Jain 2001). In hydraulics, it is also called
negative surge (Montes 1998; Chanson 2004), implying shallow waters invading
deeper waters. Its fundamental equation is thus
5.6 Simple Wave Problem: Basic Equations of Rarefaction Waves 233
From Eq. (5.83), taking the left state as reference, one can write
U þ 2c ¼ UA þ 2cA : ð5:85Þ
1 x
c¼ UA þ 2cA ; ð5:86Þ
3 t
or
1 x 2
hðx; tÞ ¼ UA þ 2cA ; ð5:87Þ
9g t
234 5 Unsteady Open Channel Flows: Basic Solutions
and
1 2x
U ðx; tÞ ¼ UA þ 2cA þ : ð5:88Þ
3 t
The SWE have simplified forms, upon neglecting specific terms of the momentum
equation, widely used in hydrology. Based on the terms retained, dynamic, diffu-
sive, and kinematic waves are defined as follows (Chow et al. 1988)
@U @U @h
þU þ g g So Sf ¼ 0: ð5:89Þ
@t @x @x |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}
kinematic wave
|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
diffusive wave
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
dynamic wave
With the actual power of computational methods, the dynamic wave model is by far
the dominant approach. Only the kinematic wave is, however, still used in
hydrological models. Given this fact and the conceptual interest of kinematic
waves, this simplified model is described here. The kinematic wave equations are,
for a rectangular channel, the continuity equation
@h @q
þ ¼ 0; ð5:90Þ
@t @x
and a simplified momentum equation stating balance of gravity and friction forces,
So Sf ¼ 0: ð5:91Þ
The kinematic wave implies that: (i) The flow is not significantly accelerated, and
(ii) the variation of flow depth with distance in a given instant of time is small with a
negligible pressure gradient. The friction slope is assumed to produce a relation of
the type
h ¼ aqb : ð5:92Þ
For example, using Manning’s equation for a wide rectangular channel, the fol-
lowing result in conformity with Eq. (5.92) is generated
5.7 Simplified Models: The Kinematic Wave 235
2 3=10
n
h¼ q3=5 : ð5:93Þ
So
@q @q
abqb1 þ ¼ 0: ð5:94Þ
@t @x
@q @q
dq ¼ dt þ dx; ð5:95Þ
@t @x
236 5 Unsteady Open Channel Flows: Basic Solutions
which is rewritten as
Dq @q @q dx
¼ þ : ð5:95Þ
Dt @t @x dt
Dq
¼ 0; ð5:96Þ
Dt
dx 1
¼ : ð5:97Þ
dt abqb1
This is the characteristic form of the kinematic wave model. It implies that the
discharge q is a constant along the paths in the (x, t) plane defined by Eq. (5.97).
These paths are forward characteristics, thereby implying that in this type of wave,
information of a perturbation cannot be transmitted forward in time and back
in space. Basically, this means that only one boundary condition upstream of the
river reach is necessary for kinematic wave routing, e.g., the inflow hydrograph
q = q(0, t). Note by differentiation of Eq. (5.92) that
dh
¼ abqb1 : ð5:98Þ
dq
Thus, the kinematic wave characteristics are given with ck as the kinematic wave
celerity (Chow et al. 1988) by
dx dq
¼ ¼ ck : ð5:99Þ
dt dh
If the inflow discharge is known at the upstream end of the reach, a specific value
q = q(0, t) of the discharge at a time t will appear at a position L at the end of the
reach (Fig. 5.21) at time s given by
ZL
dx L
s ¼ tþ ¼ tþ : ð5:100Þ
c k ð qÞ ck ½qð0; tÞ
0
Fig. 5.21 Kinematic wave routing (adapted from Chow et al. 1988)
unaltered and the wave profile simply suffers a deformation (Fig. 5.21) (Henderson
1966). This deformation implies a steepening of the tailwater face, eventually
leading to the formation of a kinematic shock (Lighthill and Whitham 1955).
However, this issue is debatable, as shown by Henderson (1966), given that the
steepening is accompanied by an increase in the importance of free surface slope
terms neglected in the kinematic wave model. These terms may become important
at the threshold of formation of a kinematic surge, given that they introduce dif-
fusion and attenuation. If they are strong enough, the kinematic surge formation
may be counterbalanced and a smooth wave profile of permanent shape traveling at
constant speed finally forms. This smooth wave solution is called monoclinal wave
(Sturm 2001; Jain 2001). In the more complex case of flow with rainfall, a source
term must be added to the continuity equation and the characteristic lines are then
curved (Chow et al. 1988; Jain 2001).
The kinematic wave celerity can be written as
dq dðUhÞ dU
ck ¼ ¼ ¼ U þh : ð5:101Þ
dh dh dh
S1=2
U¼ o
h2=3 ð5:102Þ
n
results in
dU 2 So1=2 1=3
¼ h : ð5:103Þ
dh 3 n
238 5 Unsteady Open Channel Flows: Basic Solutions
5
ck ¼ U: ð5:104Þ
3
Comparing Eq. (5.104) with the absolute celerity cd of the dynamic wave traveling
downstream
cd ¼ U þ ðghÞ1=2 ; ð5:105Þ
ck 5
3U 5 F
¼ ¼ : ð5:106Þ
cd U þ ðghÞ1=2 3 F þ 1
Lighthill and Whitham (1955) demonstrated that in a flood wave propagating over
initially uniform flow, both kinematic and dynamic waves coexist. Consider a
smooth and continuous flood wave propagating over an initially steady and uniform
stream. They argued that the main body of the flood behaves, essentially, like a
kinematic wave, whereas the wave fronts are basically governed by the dynamic
wave model. If F\1:5, then Eq. (5.106) yields ck < c, implying that the main
body of the flood wave travels slower than the front advancing at velocity
Uo + (gho)1/2. In reality, close to, but slightly above F 1, the flow is
non-hydrostatic, with undular features, a characteristic beyond the scope of the
SWE. Thus, for a subcritical stream with F\1, which is clearly the most usual case
in river flows, the main body of the wave can in fact propagate at celerity ck as
found by Lighthill and Whitham (1955), whereas the leading positive front would
propagate at speed Uo+ (gho)1/2 (Fig. 5.22) implying that there is a “dynamic wave
ahead the main kinematic wave.” They also showed that in the domain of solution
where the dynamic wave is the governing model, there is no appreciable variation
of flow depth and velocity. Of course, this is likely to be linked to the assumption of
initially uniform flow. The strongest variations in both are accomplished once the
main (kinematic) bulk of the wave arrives (Henderson 1966) (Fig. 5.22).
As previously explained, Eq. (5.44) are the non-conservative form of the SWE.
Manipulation permits to write this system as (Cunge 1975; Toro 2001)
5.8 Use of Non-conservative Form of SWE 239
Fig. 5.22 Kinematic and dynamic waves (adapted from Stoker 1957; Henderson 1966)
@ h @ Uh 0
þ ¼ : ð5:107Þ
@t U @x 1 2
2 U þ gh 0
This equation is generally in good agreement with experiments for breaking pos-
itive surges (Cunge 1975). The integral form of Eq. (5.107) yields as celerity of the
shock wave (Cunge 1975; Toro 2001)
240 5 Unsteady Open Channel Flows: Basic Solutions
1=2
2gh21
cNC ¼ : ð5:109Þ
ð h1 þ h2 Þ
This ratio is only equal to unity in the trivial case h1 = h2, e.g., for smooth solu-
tions. However, Eq. (5.110) shows in Fig. 5.23 that for h1/h2 < 2, the errors of
Eq. (5.107) predicting the celerity of a positive surge are small. Thus, the
non-conservative form of the SWE applies to compute smooth or continuous
solutions, as previously done using the method of characteristics, and for weak
shock waves. Physically, Eq. (5.109) is based on conservation of mass and energy
across a shock wave. Clearly, a shock wave is a highly dissipative phenomenon,
and, thus, a solution based on energy conservation is incorrect. For weak shocks,
however, energy dissipation is small, explaining why Eq. (5.109) is acceptable for
small surge heights. Thus, it is necessary to write the conservation laws in con-
servative form to obtain weak solutions, but, also, the conserved variables should
have a physical meaning in the problem being simulated (Cunge 1975).
A physical shock implies characteristics converging from both sides of the shock
path in the (x, t) plane (Cunge 1975; Toro 2001). It is possible to mathematically
obtain non-physical shocks due to the entropy-violating conditions, where char-
acteristics diverge (Toro 2001; Katopodes 2019). Obviously, these weak solutions
of the SWE must be discarded.
5.9 Limitations of SWE 241
Here, the control volume CV is an area of the (x, t) plane, and the flux integral is
extended along the closed contour line which defines the CV, with s as the curvi-
linear arch length. Consider the spillway depicted in Fig. 5.24. Static water at the
upstream face of the weir is considered as initial condition. Boundary conditions are
physically a constant discharge supplied at the upstream inlet, and a prescribed
tailwater flow depth obtained by regulation of a gate. After transient flow, the
steady flow profile shown in Fig. 5.24 is generated.
SWE produce a continuous steady flow solution at the weir crest, where the flow
changes from sub- to supercritical conditions. Somewhere in the tailwater, a
hydraulic jump is formed to allow for transitional flow from super- to subcritical
conditions forced with a gate. Both types of transcritical flows are therefore
obtained as solutions of SWE [Eqs. (5.111)] (numerical computations to be detailed
in Chap. 9). However, these mathematical results might not be an accurate repre-
sentation of the physical phenomena being simulated: If the fluid pressure is in
reality non-hydrostatic, then the mathematical solution of the SWE is meaningless
(see Chap. 11). Roughly, for example, the continuous flow solution of the SWE at a
weir crest is only accurate if the ratio of crest specific energy to bottom curvature is
below 0.25 (Castro-Orgaz and Chanson 2016). The hydraulic jump is correctly
predicted by the SWE only if a roller is formed (Castro-Orgaz and Hager 2009). If
the hydraulic jump is undular, say for inflow Froude numbers below 1.7 (Montes
and Chanson 1998), then the prediction of the SWE is not in conformity with
experimental observations.
Fig. 5.24 Steady transcritical flow as solution of unsteady computations using SWE
242 5 Unsteady Open Channel Flows: Basic Solutions
Fig. 5.25 Dike erosion test showing continuous weir flow surface over eroded dike and
hydraulic jump in tailwater (test at experimental flume of IAS-CSIC, Córdoba; photograph by
O. Castro-Orgaz)
The continuous flow solution at a weir crest and a moving hydraulic jump
(positive surge propagating upstream) can be observed in many unsteady envi-
ronmental flows. A prominent example is the gradual erosion of dikes (Fig. 5.25).
These flows are simulated using extended versions of the SWE accounting for the
suspended and bed-load sediment transport modes (Wu 2008) (see Chap. 10).
Limitations of predictions based on these families of models are essentially those
described above; e.g., the fluid pressure shall be hydrostatic, in addition to these
related to empirical sediment transport formulae.
There exist cases where the flood routing procedure can be accomplished by
ignoring dynamic effects, e.g., neglecting the momentum equation in the SWE and
solving the continuity equation alone. An important example involves the passage
of a flood wave across a reservoir (Fig. 5.26), given that the reservoir water surface
remains nearly horizontal. Consider here a reservoir with an overflow structure
consisting of a standard spillway crest without gates.
The continuity equation is
@A @Q
þ ¼ 0: ð5:112Þ
@t @x
Zxs
d
Qe Qs ¼ Adx: ð5:113Þ
dt
xe
Zxs
8¼ Adx; ð5:114Þ
xe
d8
¼ Qe Qs : ð5:115Þ
dt
The inflow hydrograph is typically a function of time, e.g., Qe = Qe(t), whereas the
spillway rating curve is of the form
h i
Qs ¼ Cd Lð2gE 3 Þ1=2 Cd ð2gÞ1=2 Lh3=2 ¼ CLh3=2 : ð5:116Þ
Here, E is the energy head on the weir crest, approximated to the approach flow
depth above the spillway invert h, L is the spillway width, and Cd is the discharge
coefficient. Note that the C coefficient used in hydrological computations is not
dimensionless. For a known inflow hydrograph and given spillway, Eq. (5.115)
must be numerically solved to compute the outflow hydrograph.
First, the modified Puls method (Chow et al. 1988) is presented here. It is an
implicit finite difference method which is based on the integral form of Eq. (5.115)
over a time period Δt = ti+1 − ti, that is,
244 5 Unsteady Open Channel Flows: Basic Solutions
ZSi þ 1 Zti þ 1
dS ¼ ðQe Qs Þdt: ð5:117Þ
Si ti
Here, i represents the time subscript, and S is the storage volume above the spillway
crest. Using the trapezoidal rule, Eq. (5.117) yields
Dt
Si þ 1 S i ¼ ðQe Qs Þi þ 1 þ ðQe Qs Þi ; ð5:118Þ
2
or re-arranged
2Si þ 1 2Si
þ Qsi þ 1 ¼ ðQei þ Qei þ 1 Þ þ Qsi : ð5:119Þ
Dt Dt
To apply Eq. (5.119), the inflow hydrograph Qe = Qe(t) and discharge rating curve
Qs = Qs(h) are known in advance. However, to produce the output Qs = Qs(t) as
solution of Eq. (5.119), a function relating S to the actual value of h is additionally
needed. This information results from topographic data of the reservoir by evalu-
ating the integral
Zh
Sð hÞ ¼ F ðhÞdh: ð5:120Þ
0
Here, F(h) is the plan area of the water surface at elevation h relative to the spillway
crest. The function F = F(h) is known from constant elevation contour lines in the
reservoir, taken from available planimetric information. In general, the volume
integral must be numerically solved using a low order method, thereby introducing
errors in the estimation (Fenton 1992). Once S = S(h) is available, the numerical
procedure is as follows:
1. As initial condition, the water surface elevation is assumed to be at the spillway
invert when the inflow hydrograph reaches the upstream reservoir section. Then,
both S and Qs are zero at t = 0. Of course, other scenarios are possible. The
quantity 2S/Δt − Qs is thus zero. The sum Qei + Qei+1 is known from the inflow
hydrograph for all values of i.
2. A routing interval Δt is selected. The function 2S/Δt + Qs versus Qs is con-
structed using the functions Qs = Qs(h) and S = S(h): For a given h, S and Qs are
determined and thus 2S/Δt + Qs.
3. The variable 2S/Δt + Qs at i + 1 is computed from Eq. (5.119).
4. Using the function 2S/Δt + Qs versus Qs, the outflow discharge Qsi+1 is com-
puted from the known value (2S/Δt + Qs)i+1.
5.10 Hydrologic Routing 245
5. Apply the identity 2S/Δt − Qs = (2S/Δt + Qs) − 2Qs to generate the initial data
for the next time step.
6. Go back to Step 3 if the final time is not yet reached.
An alternative to use the integral Eq. (5.117) is to directly solve Eq. (5.115),
which is a first-order ODE (Chow et al. 1988; Ayuso 1990; Fenton 1992). The
change in water storage for an infinitesimal change in surface elevation dzs is
d8 ¼ Fdzs : ð5:121Þ
Inserting Eq. (5.121) in Eq. (5.115) produces
dzs Qe Qs
¼ ; ð5:122Þ
dt F
or,
dh Qe ðtÞ Qs ðhÞ
¼ f ðt; hÞ: ð5:123Þ
dt F ð hÞ
This is the equation describing the temporal variation of the reservoir level h = h(t).
Equation (5.123) can be solved using any method to solve ODEs. If the time
derivative is discretized using a forward in time finite-difference
dh hi þ 1 hi
¼ ; ð5:124Þ
dt Dt
Qe ðti Þ Qs ðhi Þ
hi þ 1 ¼ hi þ fi Dt ¼ hi þ Dt: ð5:125Þ
F ð hi Þ
This equation has the advantage of being extremely simple to apply. Given its
explicit character, the time step Δt must be limited to assure numerical stability.
However, as demonstrated by Fenton (1992), this is not a real limitation in practical
computations. Equation (5.125) is the so-called Euler equation, a first-order
Runge–Kutta method. Runge–Kutta schemes are explicit, using an updating for-
mula of the type (see Chap. 3)
hi þ 1 ¼ hi þ f Dt; ð5:126Þ
where the average value of f is determined differently, depending on the order of the
scheme. Chow et al. (1988) proposed a third-order Runge–Kutta method, whereas
Ayuso (1990) used a fourth-order scheme. For this last option, the average slope is
246 5 Unsteady Open Channel Flows: Basic Solutions
1
f ¼ ðk1 þ 2k2 þ 2k3 þ k4 Þ; ð5:127Þ
6
with
k1 ¼ f ðti ; hi Þ;
1 1
k2 ¼ f ti þ Dt; hi þ k1 Dt ;
2 2
ð5:128Þ
1 1
k3 ¼ f ti þ Dt; hi þ k2 Dt ;
2 2
k4 ¼ f ðti þ Dt; hi þ k3 DtÞ:
Following Fenton (1992), a comparison of the modified Puls method, the Euler
method, and the fourth-order Runge–Kutta method is shown in Fig. 5.27. The
inflow hydrograph considered is plotted in the figure; note that the peak discharge is
120 m3/s and the base time about 6 h. The reservoir is prismatic with a constant
horizontal area of 50,000 m2. The spillway for this test has a width L = 100 m with
a (hydrological) discharge coefficient C = 2.037 m1/2/s. A time step Δt = 0.1 h was
adopted. The codes prepared for the computations are implemented in the files
“Euler.xls,” “Puls.xls,” and “rk4.xls,” available in Chap. 12. The code for the
fourth-order Runge–Kutta method was adapted from the Fortran code PRAVEM
available from Ayuso (1990).
Comparing the implicit (Puls) method with the explicit fourth-order Runge–
Kutta (rk4) method in Fig. 5.27a shows excellent agreement, thereby highlighting
that either of the two is a good selection. The Puls and the Euler methods are
compared in Fig. 5.27b, indicating reasonable agreement of the two. Note the small
distortion in the ascending outflow hydrograph branch predicted by Euler method,
as well as the overprediction of the peak discharge. However, despite these inac-
curacies, the simple implementation of this method supports its use at minimum for
teaching purposes (Fenton 1992).
Hager et al. (1984) experimentally determined the flood routing process in a
laboratory scale reservoir. The inflow hydrographs followed the smooth equation
Fig. 5.28 Comparison of reservoir routing solution with experiments involving smooth inflow
hydrographs with (Q*, t*, n, Cd) = a (60 l/s, 60 s, 2, 0.463), b (60 l/s, 60 s, 5, 0.46), c (150 l/s,
120 s, 10, 0.48), d (240 l/s, 60 s, 10, 0.49)
248 5 Unsteady Open Channel Flows: Basic Solutions
Fig. 5.29 Comparison of reservoir routing solution with experiments involving real inflow
hydrographs with Cd = a 0.471, b 0.475
where the normalized variables are q = Q/Q* and T = t/t*; * denotes a reference
quantity, and n is a parameter. Experiments for four tests are plotted in Fig. 5.28,
along with the values for (Q*, t*, n, Cd) in each run. The reservoir is prismatic in
elevation, with a constant plan area of F = 39.7 m2, and the spillway width is
L = 0.54 m.
The rk4 solver was run using Δt = 0.0005 h (1.8 s); the results are presented in
Fig. 5.28 superimposed with the inflow hydrograph and the measured outflows.
Note that the agreement is good, supporting the accuracy of Eq. (5.123) to describe
reservoir routing processes.
To further test the rk4 solver, an additional comparison with experiments is
made in Fig. 5.29, where the inflow hydrographs are irregular, aimed at describing
real flood events. The numerical simulation produces again a fair approximation to
the experiments, faithfully following the shape of the measured outflow discharge
curves.
The Muskingum method is a hydrologic routing technique for canals based on the
solution of the continuity equation. To match the backwater effects in a reach of a
canal or river, the storage function is composed of two terms, one representing the
backwater storage and the other representing a level (reservoir) storage, e.g. (Chow
et al. 1988)
5.10 Hydrologic Routing 249
S ¼ kxðI QÞ þ kQ : ð5:130Þ
|fflfflfflfflfflffl{zfflfflfflfflfflffl} |{z}
backwaterstorage levelstorage
Here, I is the inflow hydrograph and Q the outflow hydrograph. The parameters
k and x are empirical and must be therefore calibrated with data. Parameter x is a
weighting factor ranging from 0 to 0.5, controlling the wave attenuation, while
k has the dimension of time. It is usually estimated as the travel time of the peak
discharge. Using Eq. (5.130),
Si ¼ k½xIi þ ð1 xÞQi ;
ð5:131Þ
Si þ 1 ¼ k ½xIi þ 1 þ ð1 xÞQi þ 1 :
or,
Dt 2kx
C0 ¼ ;
2kð1 xÞ þ Dt
Dt þ 2kx
C1 ¼ ; ð5:134Þ
2kð1 xÞ þ Dt
2k ð1 xÞ Dt
C2 ¼ :
2kð1 xÞ þ Dt
References
Castro-Orgaz, O., & Chanson, H. (2016). Minimum specific energy and transcritical flow in
unsteady open channel flow. Journal of Irrigation and Drainage Engineering, 142(1),
04015030.
Castro-Orgaz, O., & Hager, W. H. (2009). Classical hydraulic jump: Basic flow features. Journal
of Hydraulic Research, 47(6), 744–754.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). New York: Springer.
Chaudhry, M. H. (2014). Applied hydraulic transients (3rd ed.). New York: Springer.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Chow, V. T., Maidment, D. R., & Mays, L. W. (1988). Applied hydrology. New York:
McGraw-Hill.
Cunge, J. A. (1975). Rapidly varying flow in power and pumping canals (Chap. 14). In K.
Mahmood & V. Yevjevich (Eds.), Unsteady flow in open channels, 2, 539–586. Fort Collins,
CO, USA: Water Resources Publications.
Cunge, J. A., Holly, F. M., & Verwey, A. (1980). Practical aspects of computational river
hydraulics. London: Pitman.
De Saint-Venant, A. B. (1870). Démonstration élémentaire de la formule de propagation d’une
onde ou d’une intumescence dans un canal prismatique; et remarques sur les propagations du
son et da la lumière sur les ressauts, ainsi que sur la distinction des rivières et des torrents
[Elementary demonstration of the formula of propagation of a wave or a disturbance in a
prismatic channel; and remarks on the propagations of sound, light, and bores, as also on the
distinction between a river and a torrent: Part 1]. Comptes Rendus de l’Académie des Sciences,
71, 186–195 (in French).
De Saint-Venant, A. B. (1871). Théorie du mouvement non permanent des eaux, avec application
aux crues des rivières et à l’introduction des marrées dans leur lit [Theory of unsteady water
movement, applied to floods in rivers and the effect of tidal flows: Part 2]. Comptes Rendus de
l’Académie des Sciences, 73, 147–154 (in French).
Favre, H. (1935). Etude théorique et expérimentale des ondes de translation dans les canaux
découverts [Theoretical and experimental study of travelling surges in open channels]. Paris,
France: Dunod (in French).
Fenton, J. D. (1992). Reservoir routing. Hydrological Sciences Journal, 37(3), 233–246.
Hager, W. H., Sinniger, R., & Regamey, J.-M. (1984). Reservoir storage equation experimentally
verified. Water Power and Dam Construction, 36(11), 44–48.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan.
Hicks, F. E., & Steffler, P. M. (1990). Finite element modelling of open channel flow. Water
Resources Engineering Report 90-6. Canada: University of Alberta.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Katopodes, N. D. (2019). Free surface flow: Computational methods. Oxford, UK:
Butterworth-Heinemann.
Khan, A. A., & Lai, W. (2014). Modeling shallow water flows using the discontinuous Galerkin
method. New York: CRC Press, Taylor and Francis.
Lai, C. (1986). Numerical modelling of unsteady open-channel flow. Advances in Hydroscience,
14, 161–333.
LeVeque, R. J. (2002). Finite volume methods for hyperbolic problems. New York: Cambridge
University Press.
Lighthill, M. J., & Whitham, G. B. (1955). On kinematic waves: Flood movement in long rivers.
Proceedings Royal Society London, A, 229, 281–345.
Liggett, J. A. (1994). Fluid mechanics. New York: McGraw-Hill.
Macdonald, I. (1995). Analysis and computation of steady open channel flow (Ph.D. thesis).
University of Reading, UK.
Montes, J. S. (1998). Hydraulics of open channel flow. Reston, VA: ASCE.
Montes, J. S., & Chanson, H. (1998). Characteristics of undular hydraulic jumps: Results and
calculations. Journal of Hydraulic Engineering, 124(2), 192–205.
References 251
6.1 Introduction
Fig. 6.1 Dam breaks of a Möhne Dam, b Eder Dam. Photographs taken from the webpage http://
www.thedambusters.org.uk/
D dx
ðU 2cÞ ¼ 0 if ¼ U c;
Dt dt ð6:1Þ
D dx
ðU þ 2cÞ ¼ 0 if ¼ U þ c:
Dt dt
Here, U is the water velocity, c = (gh)1/2 the shallow water wave celerity, g the
gravity acceleration, h the water depth, x the space coordinate, and t time.
After dam break initiation, a negative wave spreads back into the reservoir, while
a wave front of zero depth travels forward over the dry channel bed. Line OB
represents the border of the negative wave, which is always in contact with still
water where the water depth equals ho. Thus, this is a straight backward C−
characteristic; all other C− characteristics are also straight lines (Stoker 1957; Jain
2001). Its equation is
dx x
¼ Uo co ¼ ¼ ðgho Þ1=2 : ð6:2Þ
dt t
dx
¼ UF : ð6:3Þ
dt
Given that Uo = 0 (fluid at rest), and that cF = 0 (zero depth at dry front), Eq. (6.4)
yields
2co ¼ UF : ð6:5Þ
dx x
¼ ¼ 2co ¼ 2ðgho Þ1=2 : ð6:6Þ
dt t
Accordingly, the dry front propagating over dry terrain moves at double speed of the
negative wave edge. As the slope determined by Eq. (6.6) is constant, the dry front path
is a straight line passing across origin O. From the simple wave problem, the solution
obtained from Eq. (6.1) for the C− characteristics is (see Chap. 5) [Eq. (5.82)]
x
¼ Uo þ 2co 3½ghð0; sÞ1=2 : ð6:7Þ
ts
Here, h(0, s) is the flow depth at the point of intersection of the C− characteristic
with the t-axis. In the dam break problem Uo= 0, simplifying Eq. (6.7) to
x
¼ 2co 3½ghð0; sÞ1=2 : ð6:8Þ
ts
At the dam axis, x = 0. Thus, given that (t − s) > 0, one must have for all values of t
From this relation, the depth at the dam axis is a constant, given by
4 c2o 4
h¼ ¼ ho ; ð6:10Þ
9g 9
or,
2
c ¼ co : ð6:11Þ
3
The basic equation of the simple wave (Chap. 5) is written for the dam break
problem as
6.2 Dam Break Wave Under Dry Tailwater Conditions 257
Applying Eq. (6.12) to the dam axis, the corresponding velocity is a constant given by
4 2
Uþ co ¼ 2co ) U ¼ co : ð6:13Þ
3 3
1 h xi2
hðx; tÞ ¼ 2ðgho Þ1=2 : ð6:16Þ
9g t
Note also that along the C− characteristics both h and U are constant by virtue of the
simple wave features (Chap. 5). The SWE along the wet–dry interface degenerates
from the hyperbolic to the parabolic type (Dressler 1952). Given that all the C−
characteristics are straight lines passing across the origin O, the C+ lines are of
necessity curved (Stoker 1957; Jain 2001). The network of resulting characteristic
curves is depicted in Fig. 6.3.
In the dry-bed dam break problem for x < 0, the backward characteristics has a
negative slope, so the flow is subcritical (F\1; with F ¼ U=ðghÞ1=2 as the Froude
number). At the dam axis, the flow is critical with a backward characteristic
coincident with the t-axis. For x > 0, the backward characteristics has a positive
6.2 Dam Break Wave Under Dry Tailwater Conditions 259
Inserting the result for the free surface profile, Eq. (6.15), produces for the velocity
as function of time and space
2 hx i 2 hx i
UP ¼ U ðx; tÞ ¼ þ co ¼ þ ðgho Þ1=2 : ð6:18Þ
3 t 3 t
Equations (6.16) and (6.18) are the original solutions to the dry-bed dam break
problem of Ritter (1892). The local and convective accelerations from his solution
are, respectively,
@U 2x
¼ 2; ð6:19Þ
@t 3t
and
@U 4h xi
U ¼ ðgho Þ1=2 þ : ð6:20Þ
@x 9t t
@h 2h xi
g ¼ 2ðgho Þ1=2 : ð6:21Þ
@x 9t t
@U @U @h
þU þg ¼ 0: ð6:22Þ
@t @x @x
2 1=3
U q
F¼ ¼ 1 ) h ¼ hc ¼ : ð6:23Þ
ðghÞ1=2 g
Inserting Eq. (6.24) into Eq. (6.23) produces for the critical depth
This verifies that the water depth at the dam axis corresponds to the critical flow
depth. Given that the solution at the dam axis is steady, the specific energy concept
applies. There, E is a minimum given by
q2 3
E ¼ hþ ¼ hc : ð6:26Þ
2gh2 2
Inclusion of hydraulic resistance notably modifies the flow behavior in the vicinity
of the dry front,1 yet the conditions near the dam axis and the negative front are not
greatly affected.
A finite water depth is now considered in the downstream channel (Fig. 6.4a).
Ritter’s parabolic profile cannot be simply connected to the stillwater layer, given
that the velocity is then discontinuous at the assembling point B (finite at the
parabolic edge, zero at the still water). Physically, the only possibility to account for
a velocity discontinuity is the formation of a surge of finite height, where both
velocity and depth are discontinuous (Fig. 6.4b) (Stoker 1957; Henderson 1966).
The structure of a dam break wave under wet tailwater conditions implies, there-
fore, a rarefaction wave connected to a surge advancing over the still water
(Fig. 6.5).
1
The effect of hydraulic resistance on dam break waves is a complex problem which was notably
tackled using analytical developments by Dressler (1952, 1954) and Whitham (1955), and
experimentally by Schoklitsch (1917). However, no general analytical solution is yet known to the
problem. Thus, this topic will be presented within a numerical framework in Chap. 9 using finite
volume solvers of the SWE.
6.3 Dam Break Wave Under Wet Tailwater Conditions 261
Thus, we have to determine the surge front and ensemble it to Ritter’s parabolic
wave profile. The new unknowns are the water depth and velocity in the surge zone
(UB, hB), and the surge front velocity Vw (Montes 1998). At this point, it is illus-
trative to detail an alternative method to the Reynolds transport theorem (see Chap. 5)
to obtain the surge control-volume equations. The surge is a discontinuity in depth and
velocity that propagates at constant speed Vw. Therefore, its shape is not deformed as it
moves, so it is possible to transform the SWE PDEs (see Chap. 5),
262 6 Ideal Dam Break Waves
@U @F
þ ¼ 0; ð6:27Þ
@t @x
to moving axes with the surge using a Galilei coordinate transformation, where
U = (h, Uh)T and F = (Uh, gh2/2 + U2h)T are the vectors of unknowns and fluxes,
respectively. The space coordinate X in moving axes is related to the absolute
longitudinal coordinate x by X = x − Vwt. Therefore, basic calculus permits to write
the identities
@U @U @X @U
¼ ¼ ðVw Þ; ð6:28Þ
@t @X @t @X
@F @F @X @F
¼ ¼ ð þ 1Þ: ð6:29Þ
@x @X @x @X
The SWE are now transformed for a wave traveling with constant speed and shape to
@U @F @U @F
þ ¼ Vw þ ¼ 0; ð6:30Þ
@t @x @X @X
d
ðF Vw UÞ ¼ 0: ð6:31Þ
dX
These ODEs do not apply at the discontinuity of a surge. Therefore, its integral
form is
ZXd
d
ðF Vw UÞdX ¼ 0; ð6:32Þ
dX
XB
which yields the following solution as a function of the states ahead and behind the
shock
FB Fd
¼ Vw : ð6:33Þ
UB Ud
This vector equation states the so-called Rankine–Hugoniot jump conditions for a
shock wave. Therefore, the two equations to be satisfied are the mass and
momentum balances across the surge, given, respectively, from the Rankine–
Hugoniot jump conditions Eq. (6.33)
6.3 Dam Break Wave Under Wet Tailwater Conditions 263
V w ð hB hd Þ ¼ U B hB ; ð6:34Þ
2
h2B h
Vw ðUB hB Þ ¼ g þ UB2 hB g d : ð6:35Þ
2 2
hB ðVw UB Þ ¼ Vw hB ; ð6:36Þ
and
Equations (6.36) and (6.38) are identical to those found by applying Reynolds’
transport theorem in a fixed reference system (see Chap. 5). The three unknowns to
fully determine the wave structure are UB, hB, and Vw. A third equation needed to
mathematically close the system originates from conserving the Riemann invariant
along a forward characteristic C+ connecting the joint point of the parabolic and
surge portions (point B in Fig. 6.5), with a point on the negative front. The resulting
equation is given by
Its solution to find the root hB is easily implemented using the Newton–Raphson
method as follows. Equation (6.40) is rewritten for convenience as
1=2 h i
g ð hB þ hd Þ
f ðhB Þ ðhB hd Þ þ 2 ðghB Þ1=2 ðgho Þ1=2 ¼ 0: ð6:41Þ
2 hd hB
264 6 Ideal Dam Break Waves
fk
hkBþ 1 ¼ hkB : ð6:42Þ
ðdf =dhB Þk
Once hB is computed, Vw follows from the momentum balance applying Eq. (6.38),
and UB is given from Eq. (6.39) stating the Riemann invariant along the C+
characteristic by
or,
1 1 hd 1=2 1
1þ ¼ 0: ð6:47Þ
2 hd =hB hB ð1 hd =hB Þ
The solution of this implicit equation produces hd/hB 0.3105. Thus, the limiting
ratio of water depths is hd/ho = 0:3105 ð4=9Þ ¼ 0:138. For hd > 0.138ho, the wave
structure is as depicted in Fig. 6.7, with subcritical flow throughout.
The solution for the ideal dry and wet dam break waves described above is
implemented in a source code available in the sheet DamBreakAnalytical.xls
(Chap. 12). An example using this code is presented here, applying it to simulate
the ideal dam break of the Möhne Dam. The maximum water depth in the reservoir
was ho = 32 m (Kirschmer 1949). It was observed that a surge of about 10 m
height propagated in the tailwater. Simulations were conducted assuming hd =
1.5 m. The numerical solution gave hB = 9.7 m, UB = 15.93 m/s, and
Vw = 18.84 m/s. The computed free surface and velocity profiles 1 h after the dam
break are plotted in Fig. 6.8. Note that the surge height is close to that observed.
The relative wave celerity is cw = Vw− UB = 2.9 m/s, which is also in fair agree-
ment with estimations available; an average value of 2.88 m/s is quoted by
Kirschmer (1949). This computation gives an impression of the tremendous mag-
nitude of the flow variables after breaking of a real dam.
Fig. 6.8 Ideal dam break wave of Möhne Dam for 1.5 m in the tailwater, after 1 h of routing
6.4 Computational Examples 267
Fig. 6.9 Ideal dam break wave of Möhne Dam for dry tailwater after 1 h of routing
The same computation was repeated for the dry-bed condition (hd = 0), resulting
in the profiles plotted in Fig. 6.9. As observed, the increase of velocity at the
positive front is considerable. Comparing these two sets of simulations for the
Möhne Dam results in Vw/(2co) = 18.84/35.43 = 0.531. This shows the tremen-
dously decelerating effect of wetting the tailwater. A plot of Vw/(2co) as a function
of the depth ratio hd/ho is presented in Fig. 6.10, to show with more generality the
behavior of the front celerity as a function of tailwater wetting conditions. Note that
the deceleration effect occurs fast just after a minimal wetting of the tailwater (see
detail in Fig. 6.10). The positive front velocity ratio then becomes nearly constant
with values around Vw/(2co) = 0.47, and then slightly increases when hd/ho raises
above 0.34.
The quality of Ritter’s solution to predict dry-bed dam break waves is examined
in Fig. 6.11, where the free surface predictions using Eq. (6.16) are compared with
the experimental data of Ozmen-Cagatay and Kocaman (2010) at different nor-
malized times T = t(g/ho)1/2 starting at abrupt gate removal. The shape of the dam
break wave curves of Fig. 6.11 is similar to those previously measured by Dressler
(1954) and by Lauber and Hager (1998). The upstream water depth in the exper-
iments by Ozmen-Cagatay and Kocaman (2010) was ho = 0.25 m. Note that the
parabolic shape predicted by Ritter is in fair agreement with observations for this
dataset. For a more general prediction of dry-bed dam break waves, non-hydrostatic
and friction effects have to be accounted for. Castro-Orgaz and Chanson (2017)
found that the negative wave celerity and the wave profile are accurately predicted
if non-hydrostatic effects are considered in a depth-averaged model, whereas fric-
tion effects for laboratory data are confined to the narrow front tip which behaves
essentially like a diffusive wave.
Computations for wet-bed dam break waves are compared with laboratory data
for hd/ho < 0.138 and hd/ho > 0.138 in Figs. 6.12 and 6.13, respectively. During
wave initiation, the wave profile is affected by the non-hydrostatic flow condition, a
feature beyond the capabilities of the SWE, and, thus, agreement with experiments
is not expected during this process (see Chap. 11). For a wet-bed dam break wave
of shallow tailwater (Fig. 6.12), supercritical flow occurs in a part of the wave
profile. As shown by experiments, the turbulence and wave breaking are strong in
these flows, producing an irregular free surface that is not in precise agreement with
the theoretical prediction. However, the overall position of the surge front is rea-
sonably predicted.
For the subcritical dam break wave (hd/ho > 0.138), turbulence effects are less
stringent, and once the wave breaks, the prediction of the SWE is in excellent
agreement with experiments (see Fig. 6.13 at T = 6.51 and 6.89).
6.4 Computational Examples 269
References
Castro-Orgaz, O., & Chanson, H. (2017). Ritter’s dry-bed dam-break flows: Positive and negative
wave dynamics. Environmental Fluid Mechanics, 17(4), 665–694.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Chow, V. T. (1959). Open channel hydraulics. New York: McGraw-Hill.
Dressler, R. F. (1952). Hydraulic resistance effect upon the dambreak functions. Journal of
Research, National Bureau of Standards, 49(3), 217–225.
Dressler, R. (1954). Comparison of theories and experiments for the hydraulic dam-break wave.
Proceedings of the International Association Scientific Hydrology, Assemblée Générale, 3(38),
319–328, Rome, Italy.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan Co.
Hoffman, J. D. (2001). Numerical methods for engineers and scientists (2nd ed.). New York:
Marcel Dekker.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Kirschmer, O. (1949). Zerstörung von Talsperren und Dämmen [Destruction of dams and dikes].
Schweizerische Bauzeitung, 67(20), 277–281 (in German).
Lauber, G., & Hager, W. H. (1998). Experiments to dambreak wave: Horizontal channel. Journal
of Hydraulic Research, 36(3), 291–307.
Montes, J. S. (1998). Hydraulics of open channel flow. Reston VA: ASCE.
Ozmen-Cagatay, H., & Kocaman, S. (2010). Dam-break flows during initial stage using SWE and
RANS approaches. Journal of Hydraulic Research, 48(5), 603–611.
Ré, R. (1946). Étude du lacher instantané d’une retenue d’eau dans un canal par la méthode
graphique [Study of instantaneous release of water in a channel using the graphical method].
La Houille Blanche, 1(3), 181–188 (in French).
Ritter, A. (1892). Die Fortpflanzung von Wasserwellen [Propagation of water waves]. Zeitschrift
Verein Deutscher Ingenieure, 36(2), 947–954 (in German).
Schoklitsch, A. (1917). Über Dammbruchwellen [On dam break waves]. Kaiserliche Akademie
der Wissenschaften, Wien, Mathematisch-Naturwissenschaftliche Klasse. Sitzungberichte IIa,
126, 1489–1514 (in German).
Stoker, J. J. (1957). Water waves: The mathematical theories with applications. New York: Wiley.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. Singapore: Wiley.
Toro, E. F. (2009). Riemann solvers and numerical methods for fluid dynamics. London: Springer.
Whitham, G. B. (1955). The effects of hydraulic resistance in the dam-break problem. Proceedings
of the Royal Society of London, A, 227, 399–407.
Chapter 7
Finite Difference Methods
7.1 Introduction
The SWE are a system of two nonlinear hyperbolic PDEs that must be numerically
solved to describe the time evolution of the fluid velocity and water depth in the
entire computational domain. Finite-difference methods to obtain approximate
numerical solutions are described in this chapter. First, basic numerical aspects are
presented. The implementation of boundary conditions for continuous and dis-
continuous flows is then discussed, and various schemes widely used are explained
in detail. The performance of these schemes is assessed using analytical solutions
and experimental data for selected test cases.
Consistency: This refers to the relation between the algebraic equation produced by
the FD scheme and the original differential equation. Consistency is reached if the
FD equation approaches the original differential equation as Dx and Dt tend to zero.
Consistency is needed to have a convergent solution, but it is not a sufficient
condition.
Stability: A numerical scheme is stable if the solution remains bounded as it evolves
in time. An unstable solution becomes unbounded due to the accumulation of errors
and may result in a crashing of computations.
The SWE for a rectangular cross section, written in conservative form, are
@U @F
þ ¼ S: ð7:1Þ
@t @x
Here, U is the vector of unknowns, F the flux vector, and S the source term vector,
with h as the water depth, U the fluid velocity, zb the bed elevation, x the longi-
tudinal coordinate, t the time, g the gravity acceleration, and Cf as a friction
coefficient, given by
h Uh 0
U¼ ; F¼ ; S¼ : ð7:2Þ
Uh U 2 h þ 12 gh2 gh @z
@x Cf U jU j
b
To compute the variables U(x, t) and h(x, t), a discretization of the governing
equations is required. The partial derivatives (or integrals) in the governing equa-
tions are then replaced by discrete numbers. Discretization of the partial differential
equations, Eq. (7.1), leads to finite difference methods (FDM). In turn, the dis-
cretization of integral equations as
ZZ I ZZ
@U
dxdt þ ðF nÞds ¼ Sdxdt; ð7:3Þ
@t
CV CV
leads to the so-called finite volume methods (FVM) (Anderson 1995; Hoffman
2001; Toro 2009), with CV as the control volume in the (x, t) plane, n as the unit
vector normal to the closed contour of CV, and s as the curvilinear coordinate. Here,
FDM are considered, where the PDEs are approximated by algebraic relations
obtained by replacing the partial derivatives by finite-differences.
7.2 Basic Numerical Aspects 275
Consider a finite-difference grid in the x-t plane (Fig. 7.1), where nodes are
defined by uniform steps Dx and Dt in the x- and t-directions, respectively. The
objective is to replace the original PDEs, valid at any point on the x-t plane, by
algebraic expressions that apply only to the grid points (i, k) of the mesh, with i and
k as the node indices in the x- and t-directions, respectively.
Finite-difference representations of derivatives are based on Taylor series
expansions (Sturm 2001; Chaudhry 2008). For example, for vector F, a Taylor
expansion in the x-direction at point i + 1 and time level k reads (Hoffman 2001)
k k 3 k
@F @ 2 F ðDxÞ2 @ F ðDxÞ3
Fkiþ 1 ¼ Fki þ Dx þ þ þ : ð7:4Þ
@x i @x2 i 2 @x3 i 6
Solving for the first derivative of F from the former expression results in
k 2 k 3 k
@F Fkiþ 1 Fki @ F Dx @ F ðDxÞ2
¼ þ : ð7:5Þ
@x i Dx
|fflfflfflfflfflffl{zfflfflfflfflfflffl} @x2 i 2 @x3 i 6
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Finite difference approximation Truncation error
If the first derivative in Eq. (7.5) is taken as the finite-difference approximation, the
remaining part of the equation is the so-called truncation error, thus
k
@F Fk Fki
iþ1 : ð7:6Þ
@x i Dx
k
@F Fkiþ 1 Fki
¼ þ OðDxÞ: ð7:7Þ
@x i Dx
From similar arguments, one finds for the so-called backward finite-difference
k
@F Fk Fki1
¼ i þ OðDxÞ: ð7:9Þ
@x i Dx
@f @f @2f
þ u ¼ D 2 : ð7:12Þ
@t @x
|{z} @xffl}
|fflffl{zffl
convective term diffusion term
@f @f
þ u ¼ 0: ð7:13Þ
@t @x
While the obvious choice for discretization of the time derivative to advance in
time starting at known time level is the forward finite-difference
@f fik þ 1 fik
¼ þ O(DtÞ; ð7:14Þ
@t Dt
the best choice for the spatial gradient is not so evident. The following options are
mathematically possible:
Forward finite-difference
@f fikþ 1 fik
¼ þ O(DxÞ; ð7:15Þ
@x Dx
Backward finite-difference
@f fik fi1
k
¼ þ O(DxÞ; ð7:16Þ
@x Dx
Central finite-difference
@f fikþ 1 fi1
k
¼ þ O(Dx2 Þ: ð7:17Þ
@x 2Dx
Dt þu Dx ¼ 0 if u [ 0,
fik þ 1 fik f k 1 fik
ð7:18Þ
Dt þ u i þDx ¼0 if u \0.
This scheme, due to Courant et al. (1952), is conditionally stable. Other choices
based on central [Eq. (7.17)], or one-sided finite-differences in the same direction as
u (down-winding schemes), are unconditionally unstable (Roache 1972; Anderson
1995). A useful one-sided scheme is only obtained by accounting for the sign of the
278 7 Finite Difference Methods
wave propagation speed u. This is the utility of up-winding. The centered scheme
based on Eq. (7.17) is not good, given the unstable behavior. Note that the upwind
differences are related to the central difference by
fik fi1
k fikþ 1 fi1
k Dx fikþ 1 2fik þ fi1
k
Dx ¼ 2Dx 2 ðDxÞ2
if u [ 0,
fikþ 1 fik fikþ 1 fi1
k fikþ 1 2fik þ fi1 k ð7:19Þ
Dx ¼ 2Dx þ Dx
2 ðDxÞ2
if u \0.
Therefore, upwind methods introduce numerical diffusion into the centered scheme
to make it stable. The diffusion is numerical with D = Dx/2.
Returning to the SWE, a possible model for the discretization of Eq. (7.1)
written without source terms,
@U @F
þ ¼ 0; ð7:20Þ
@t @x
approximate) the discrete form of the integral conservation laws (to be derived for
the FVM in Chap. 9). Therefore, some finite-difference schemes approximating the
original PDEs as DU/Dt + DF/Dx = S are interpreted as particular discretizations of
the integral Eq. (7.3). In this chapter, only shock-capturing finite-difference
schemes are considered.
Two typical numerical errors appear in shock-capturing schemes. First-order
finite-difference schemes introduce numerical diffusion, smearing the shock front
(Fig. 7.2a). Second-order schemes introduce numerical dispersion errors. These
produce high-frequency oscillations near sharp gradients, as close to a shock
(Fig. 7.2b). In general, the second-order schemes are preferred in practice, given the
sharp resolution of shocks. The dispersion errors are locally controlled by adding
limited numerical diffusion.
Some finite-difference schemes mimic the integral form of the conservation laws
and thus produce shocks in the solution, as explained. However, not all
shock-capturing schemes produce physical shocks in the solution. To reproduce a
physical shock in the solution, the following entropy condition must be satisfied
(Lax 1954; Lax and Wendroff 1960; Toro 2001)
Here, k(U) = U + (gh)1/2 are the eigenvalues associated with the C+ characteristics
for the right-going shock depicted in Fig. 7.3, and L and R denote states at the left
and right sides of the shock displacing with velocity S. Equation (7.22) implies that
the characteristic curves converge to the path of the shock in the (x, t) plane
(Fig. 7.3). If Eq. (7.22) is not satisfied, the finite-difference scheme may produce a
non-physical jump (Bhallamudi 2002). Upwind models for the SWE use the
eigenvalues to detect the direction of signal propagation and thus the adequate
finite-difference representation of spatial gradients. At a critical point [U = (gh)1/2],
the eigenvalue corresponding to the backward characteristic vanishes. Then, some
280 7 Finite Difference Methods
In Eq. (7.21), the vector U at the new time level k + 1 is computed based on data at
time level k, where all information is known. It permits to evaluate F at all nodes of
the mesh. Values of U at each node i for time t + Dt are therefore explicitly solved.
This type of model is referred to as explicit. However, it is possible to evaluate F at
time level k + 1, resulting in
7.2 Basic Numerical Aspects 281
@U @FðUÞ
þ ¼ 0; ð7:24Þ
@t @x
with
! !
h U1 Uh U2
U¼ ¼ ; F¼ ¼ U22 : ð7:25Þ
Uh U2 U 2 h þ 12 gh2 U1 þ 12 gU12
Thus, Eq. (7.23) cannot be directly solved for U at the new time level k + 1. Rather,
it is an implicit equation to be solved numerically to compute the vector U at time
level k + 1. The equation at a node is thus
k þ 1
Uki þ 1 Uki Fk þ 1 Uk þ 1 Fki1
þ1
Ui1
þ iþ1 iþ1 ¼ 0: ð7:26Þ
Dt 2Dx
This type of model is said to be implicit. The implicit method results in a system of
nonlinear implicit equations, to be simultaneously solved for all nodes. They are
unconditionally stable (Chaudhry 2008). The selection of the type of scheme,
implicit or explicit, is by no means a rigid decision (Anderson 1995). The implicit
methods are applied to large time steps, given the unconditional stability. However,
the solution at each time level involves a more complex numerical method, which
obviously implies more coding work. Additionally, even if computations being
stable for any time step are an advantage, the use of a large step may deteriorate the
quality of the computed transient wave solution due to its contamination by trun-
cation errors. The explicit methods involve simple computations for each new time
level, with the corresponding simplicity of coding. However, to be stable, the time
step cannot be arbitrary. Rather, a physical constraint must be added, as explained
below, to limit the largest permissible time step. Due to coding simplicity, and if the
main interest is the computation of an accurate wave pattern which is likely to
evolve fast, as in a dam break wave, explicit schemes are a good practical choice
(Zoppou and Roberts 2003). In this chapter, the attention is thus restricted to these
techniques. To implement implicit schemes, see Sturm (2001) or Chaudhry (2008).
In an explicit method, the time step Dt for a size of cell Dx is limited to the time
needed by perturbations originating at adjacent nodes to propagate and interact.
Consider a node i, surrounded by nodes i − 1 and i + 1 in a finite-difference mesh
(Fig. 7.5). Assume that the flow is subcritical in the entire computational domain
and that Δx is fixed. If Δx is small, it is permissible to assume that the forward and
backward characteristics originating at nodes i − 1 and i + 1, respectively, are
282 7 Finite Difference Methods
Dx
ðDtÞmax ¼ : ð7:27Þ
Uc
Consider that the time step Δt selected for computations in our finite-difference mesh
is less than this value, resulting in point d at the new time level. The points on the
x-axis at the actual time level k influencing this node lie between i − 1 and i + 1.
Computationally, the analytical domain of dependence defined by the characteristics
at point d (i, k + 1) (shaded area in Fig. 7.5) thus lies inside the numerical domain of
dependence (triangle i − 1, d, i + 1) produced by the mesh. In this case, computations
are stable (Anderson 1995).
Consider now that the time step selected in the finite-difference mesh is greater
than the maximum value fixed by the numerical domain of dependence of point P
(Fig. 7.6). Therefore, the computational point d at the new time level k + 1 is above
point P. The points on the x-axis at the actual time level k influencing this node lie
outside the interval (i − 1, i + 1). The analytical domain of dependence at point d,
triggering the state of node i at the new time level k + 1, then lies outside the
numerical domain of dependence. This implies that at the new time level our
finite-difference model is demanding information about the propagation of pertur-
bations to nodes located outside the computational portion (i − 1, i + 1). This
information is not actually arriving at node d, so computations will become
unstable. The production of computations is thus forced at time Dt after the meeting
of perturbations originating from nodes i − 1 and i + 1, occurring at time (Dt)max.
7.2 Basic Numerical Aspects 283
As a consequence, the “computations arrive late in time,” by which the model loses
the information generated at nodes i − 1 and i + 1 “in the immediate past.” That is,
we demand our model to do computations at a time when all the relevant infor-
mation regarding perturbation propagation already passed through this spatial
location just a moment ago. From this, reasoning originates the stability condition
for explicit methods, stating that the Courant–Friedrichs–Lewy number CFL must
satisfy (Courant et al. 1928)
Dt
CFL ¼ ðU cÞ 1: ð7:28Þ
Dx
It assures that the analytical domain of dependence at a node lies within the
numerical domain of dependence (Anderson 1995). In other words, let a time step
be small enough to advance in time “catching” the information of perturbations
before it leaves a portion of the computational domain. From a practical viewpoint,
a common Δt for all nodes to “jump” to the new time level k + 1 is determined.
Therefore, the smallest Δt possible is selected, that is (Toro 2001; Chaudhry 2008),
CFL Dx
Dt ¼ : ð7:29Þ
maxðjUi j þ ci Þ
For practical computations, CFL ¼ 0:9 is typically used (Toro 2001; Chaudhry
2008).
284 7 Finite Difference Methods
To compute the transient wave solution of the SWE, it is necessary to know the
values of depth and velocity in all the spatial nodes of the mesh. These values
cannot be arbitrary: If these data are not compatible with the SWE, then spurious
waves will be generated in the system after initiation of the transient computations,
masking the real waves. Thus, physically realistic initial conditions must be
determined. Typically, one of the two following options is possible:
System initially static: The free surface is horizontal, and the fluid velocity is zero in
all nodes. Dry areas may be included in the computational domain.
System initially under steady flow: Depth and velocity must be determined using the
steady-state techniques (Chaps. 2–4). The solution involves the numerical solution
of the ODE for gradually varied or continuous steady flows (Henderson 1966; Jain
2001)
dh @z b
Sf
¼ @x q2 ; ð7:30Þ
dx 1 3 gh
where h is water depth, q unit discharge, zb bed elevation, and Sf the friction slope.
For rapidly varied or discontinuous steady flows (hydraulic jumps), Bélanger’s
equation is used (Henderson 1966; Jain 2001)
h2 1 h 1=2 i
¼ 1 þ 8F21 1 : ð7:31Þ
h1 2
Here, h1 and h2 are the sequent depths and F1 ¼ ðq2 =gh31 Þ1=2 is the inflow Froude
number at the supercritical section.
286 7 Finite Difference Methods
The flow state at the up- or downstream ends of the reach determines the nature of
the boundary conditions required there. For subcritical flow, one boundary condi-
tion is needed at each end of the reach, either the flow depth or the discharge or a
function relating both (rating curve). In continuous flows, the SWE in characteristic
form are applied to determine the other unknown flow variable there. For a rect-
angular cross section, and with So as the bottom slope, the SWE in characteristic
form are (Jain 2001; Chaudhry 2008)
D dx
ðU 2cÞ ¼ g So Sf if ¼ U c; ð7:32Þ
Dt dt
D dx
ðU þ 2cÞ ¼ g So Sf if ¼ U þ c; ð7:33Þ
Dt dt
or,
DU g Dh dx
¼ g S o Sf if ¼ U c; ð7:34Þ
Dt c Dt dt
DU g Dh dx
þ ¼ g So Sf if ¼ U þ c: ð7:35Þ
Dt c Dt dt
If the flow at a boundary section is discontinuous, Eqs. (7.34) and (7.35) cannot be
used (Terzidis and Strelkoff 1970; García-Navarro and Saviron 1992). Then, the
Rankine–Hugoniot jump conditions stating the mass and the momentum conser-
vation equations across a shock wave apply, namely [Eqs. (5.17) and (5.35)]
U1 h1 U2 h2
Vw ¼ ; ð7:40Þ
h 1 h2
1=2
gh1
Vw ¼ U2 þ ð h1 þ h2 Þ : ð7:41Þ
2h2
Here, Vw is the absolute surge velocity and subscripts 1 and 2 indicate conditions
up- and downstream of the shock front. Equations (7.40)–(7.41) contain five
variables (h1, h2, U1, U2, and Vw), so they apply to produce boundary conditions
provided that three are known. If a flow starts at i = 1 as discontinuous supercritical
flow due to a sudden (known) increase of discharge, the equations can be used to
generate the value of the unknown water depth there if the tailwater conditions (h2,
U2) are given.
The forward in time, centered in space scheme is possibly the simplest FD scheme,
but, as demonstrated by Liggett and Cunge (1975), it is unstable. It is therefore
called “unstable scheme.” The fluxes F are evaluated at the time level k where the
vector U is completely known, using central finite differences. The conservative
formula is then
Dt k
Uki þ 1 ¼ Uki F Fki1 þ Ski Dt: ð7:42Þ
2Dx i þ 1
This model appears to work under extremely low values of CFL in combination
with artificial viscosity to dampen oscillations. However, the scheme is of no
practical value.
7.4 Explicit Schemes 289
Peter David Lax was born on May 1, 1926, in Budapest, Hungary. His
Jewish parents both were physicists. They left Hungary in 1941, traveling via
Lisbon to New York, NY, USA. He started his studies in mathematics in
1944 at New York University (NYU). During World War II, the army sent
him to Texas A&M University, then to the Oak Ridge National Laboratory,
and soon afterward to the Manhattan Project at Los Alamos, NM, where he
began working as a calculator operator, but eventually moved on to
higher-level mathematics. After World War II, he remained with the army at
Los Alamos for another year, while taking courses at the University of New
Mexico, and then studied at Stanford University for a semester with the
Hungarian mathematicians Szegő and Pólya. He returned to NYU in 1946,
obtaining his graduation and earning the Ph.D. title in 1949 under the
supervision of Kurt O. Friedrichs (1901–1982) .
Lax made remarkable contributions early in his career continuing to
produce research, which changed the direction of many areas of mathematics.
In 1957, he published the important paper Asymptotic solutions of oscillating
initial value problems, laying the beginnings of the theory of Fourier integral
operators. Asked what was so novel about the viewpoint that made the ideas
able to enjoy such wide application, he replied: “It is a micro-local description
of what is going on. It combines looking at the problem in the large and in the
small. The numerical implementation of the micro-local point of view is by
wavelets and similar approaches, which are powerful numerically.” He
thrived in the Courant Institute of Mathematical Sciences, NYU, where
applied mathematics was studied alongside relevant pure mathematics in an
exciting mix of ideas, which led to great progress. He was appointed Institute
Director in 1972, continuing until 1980. It was a particularly difficult time to
take on this role since NYU had just closed down their School of
Engineering, moving the mathematicians from there into the Courant
Institute. This produced friction when these people wanted to set up their own
computing department while a new Computer Science Department had just
been founded. He succeeded in ensuring that there were not two rival
290 7 Finite Difference Methods
departments in the Institute, but the politics involved was difficult. His most
important works include Nonlinear partial differential equations in applied
science (1983), Hyperbolic systems of conservation laws and the mathe-
matical theory of shock waves (1987), or his Hyperbolic partial differential
equations (2006). He has received numerous awards and honors for his
mathematical researches and his advances in the field of computational
mathematics.
It is conditionally stable, and the time step Dt is restricted by the CFL condition
(Hirsch 1988, 1990; Hoffman 2001). The Lax diffusive method depends on the
value of CFL, as highlighted below. An FD scheme is said to be consistent if the
finite-difference form tends to the original PDE if both Dx and Dt ! 0. The fol-
lowing space and time Taylor expansions around the node (i, k) are formed
k 2 k
@U @ U ðDtÞ2
Uki þ 1 ¼ Uki þ Dt þ þ ; ð7:44Þ
@t i @t2 i 2
k 2 k
@U @ U ðDxÞ2
Ukiþ 1 ¼ Uki þ Dx þ þ ; ð7:45Þ
@x i @x2 i 2
k 2 k
@U @ U ðDxÞ2
Uki1 ¼ Uki Dx þ þ ; ð7:46Þ
@x i @x2 i 2
k 2 k
@F @ F ðDxÞ2
Fkiþ 1 ¼ Fki þ Dx þ þ ; ð7:47Þ
@x i @x2 i 2
k k
@F @ 2 F ðDxÞ2
Fki1 ¼ Fki Dx þ þ ð7:48Þ
@x i @x2 i 2
7.4 Explicit Schemes 291
or
@U @F 1 ðDxÞ2 @ 2 U 1 @ 2 U
þ ¼ Dt þ S: ð7:50Þ
@t @x 2 Dt @x2 4 @t2
@U @F 1 @ 2 U
þ ¼ D þ S: ð7:51Þ
@t @x 2 @x2
This is the equation modeled by the Lax scheme, where D is a numerical mesh
diffusion-like coefficient given by
ðDxÞ2
D¼ : ð7:52Þ
Dt
The Lax scheme is thus not consistent. Therefore, for a fixed value of Dx, if the
value of CFL is reduced, computations are performed with a smaller value of
Dt. Accordingly, D increases and the solution experiences greater diffusion. This
diffusive scheme has mesh-dependent features, therefore.
Dt k
Upred ¼ Uki Fi þ 1 Fki þ Ski Dt: ð7:53Þ
i
Dx
This step gives an estimation of the flow conditions at the new time level k + 1,
denoted as predicted values. In the corrector step, the fluxes F are evaluated with
the estimated vector U at time level k + 1, using backward finite-differences,
resulting in
292 7 Finite Difference Methods
Dt pred
Ucorr ¼ Uki Fi Fpred pred
i1 þ Si Dt: ð7:54Þ
i
Dx
The final value is taken as average of estimations in predictor and corrector steps,
e.g.,
Upred þ Ucorr
Uki þ 1 Uave
i ¼ i i
: ð7:55Þ
2
@U ðDxÞ2 @ 2 U
¼e : ð7:56Þ
@t Dt @x2
The method by Jameson et al. (1981) follows this idea, and it is simple to imple-
ment. The variant of Eq. (7.57) used is
ave
Uki þ 1 ¼ Uave
i þ ei þ 1=2 Ui þ 1 Ui
ave
ei1=2 Uave
i Ui1 ;
ave
ð7:58Þ
where
jhi þ 1 hi þ hi1 j
ei ¼ : ð7:60Þ
jhi þ 1 j jhi j þ jhi1 j
The “total variation diminishing (TVD)” schemes are of high order, thereby pro-
ducing a sharp resolution of shocks, but introduce a bounding of the solution to
avoid spurious oscillations. The total variation TV of a function f is defined as
(Hirsch 1988, 1990)
294 7 Finite Difference Methods
X
1
TV ð f Þ¼ jfi þ 1 fi j; ð7:61Þ
i¼1
Dt k
Uki þ 1 ¼ Uave
i þ Di þ 1=2 Dki1=2 ; ð7:63Þ
Dx
The various terms in Eq. (7.64) are systematically presented below for imple-
mentation. First, the average values of velocity and celerity are determined from
pffiffiffiffi pffiffiffiffiffiffiffiffiffi
Ui hi þ Ui þ 1 hi þ 1 hi þ hi þ 1 1=2
Ui þ 1=2 ¼ pffiffiffiffi pffiffiffiffiffiffiffiffiffi ; ci þ 1=2 ¼ g : ð7:65Þ
hi þ hi þ 1 2
k1;2
i ¼ U i ci ; k1;2
i þ 1=2 ¼ Ui þ 1=2 ci þ 1=2 : ð7:66Þ
8
< k1;2 1;2 1;2
ki þ 1=2 ei þ 1=2
i þ 1=2 if
w1;2
i þ 1=2 ¼ ; ð7:68Þ
: e1;2 else
i þ 1=2
where
e1;2
i þ 1=2 ¼ max 0; k 1;2
i þ 1=2 k 1;2 1;2
i ; k iþ1 k 1;2
i þ 1=2 : ð7:69Þ
The jump projections in the values of U between nodes i and i + 1 onto the
eigenvectors e1;2
i þ 1=2 are
1 h i
a1;2
i þ 1=2 ¼ k2;1
i þ 1=2 ðhi þ 1 hi Þ ðqi þ 1 qi Þ : ð7:70Þ
2c1i þ 1=2
The function / is a flux limiter taken here as (Hseng and Chu 2000)
2
ri1;2
þ 1=2 þ r 1;2
i þ 1=2
/1;2
i þ 1=2 ¼ 2 ; ð7:71Þ
1 þ ri1;2
þ 1=2
where
a1;2
i þ 1=2s
ri1;2
þ 1=2 ¼ ; ð7:72Þ
a1;2
i þ 1=2
and
s ¼ sign k1;2
i þ 1=2 : ð7:73Þ
In upwind methods, the discretization of the flux gradient @F=@x is sensitive to the
sign of the eigenvalues. This aims at capturing numerically the propagation of
perturbations by correctly identifying transmission up- or downstream. To illustrate
concepts, consider subcritical flow (Fig. 7.8). For perturbations propagating
296 7 Finite Difference Methods
downstream, the flux contribution is denoted by F+, whereas for those propagating
upstream the associated flux is denoted by F−. These are linked to the eigenvalues
U + (gh)1/2 and U–(gh)1/2, respectively. The flux F is then split as (Hirsch 1988,
1990)
F ¼ F þ þ F : ð7:74Þ
@U @F þ @F
þ þ ¼ 0: ð7:75Þ
@t @x @x
Dt þ Dt
Uki þ 1 Uki þ þ
Fi Fi1 þ Fi þ 1 F ¼ 0: ð7:76Þ
Dx Dx i
1
For simplicity's sake, source terms are not considered here. It is possible to treat S using an
upwind method (Toro 2001).
7.4 Explicit Schemes 297
F ¼ GU: ð7:77Þ
where
1 1
P¼ U þ pcffiffi2 U pcffiffi2 ; ð7:79Þ
and
pffiffiffi pffiffiffi
1 1 c Upffiffi2ffi p2ffiffiffi :
P ¼ ð7:80Þ
2c cþU 2 2
K ¼ K þ þ K ; ð7:82Þ
where
þ k1þ 0
K ¼ ; ð7:83Þ
0 k2þ
and
k 0
K ¼ 1
: ð7:84Þ
0 k2
298 7 Finite Difference Methods
In these matrices, the positive and negative components of the eigenvalues are
defined by the expressions
þ k1;2 þ k1;2 k1;2 k1;2
k1;2 ¼ ; k1;2 ¼ : ð7:85Þ
2 2
jKj ¼ K þ K ; ð7:86Þ
where
0
1
U þ pcffiffi2 0
jK j ¼ @ A: ð7:87Þ
cffiffi
0 U p
2
and
F þ jFj
Fþ ¼ ; ð7:92Þ
2
F jF j
F ¼ : ð7:93Þ
2
where
pffiffiffi
2 c c
c c
G11 ¼ U þ pffiffiffi U pffiffiffi U pffiffiffi U þ pffiffiffi ; ð7:95Þ
2c 2 2 2 2
p ffiffiffi
2 c c
G12 ¼ U þ pffiffiffi þ U pffiffiffi ; ð7:96Þ
2c 2 2
pffiffiffi
2 c c c
G21 ¼ U þ pffiffiffi U pffiffiffi U þ pffiffiffi
2c 2 2 2
ð7:97Þ
c c c
U pffiffiffi U pffiffiffi U þ pffiffiffi ;
2 2 2
pffiffiffi
2 c c c c
G22 ¼
U þ pffiffiffi U þ pffiffiffi þ U pffiffiffi U pffiffiffi : ð7:98Þ
2c 2 2 2 2
Dt F þ jFj F þ jFj
Uki þ 1 Uki þ
Dx 2 2
i
i1 ð7:99Þ
Dt F jFj F jFj
þ ¼ 0;
Dx 2 iþ1 2 i
or
Dt k Dt k
Uki þ 1 ¼ Uki Fi þ 1 Fki1 þ jFji þ 1 2jFjki þ jFjki1 : ð7:100Þ
2Dx ffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl 2Dx
unstable scheme diffusive term
300 7 Finite Difference Methods
Note that the upwind scheme introduces a diffusive term into the updating equation
of the unstable scheme [see Eq. (7.18) for the advection of a scalar]. The method
presented here is called flux splitting.
Consider a subcritical dam break wave of depth ratio hd/hu = 0.4 and upstream
water depth hu = 1 m. Simulations conducted with the unstable scheme using
CFL ¼ 0:1 are displayed at time t = 2 s in Fig. 7.9a, b, and compared with the
analytical solution of Stoker (1957). The numerical model is implemented in the
code available on the file “Dambreak_Unstable.xls” (Chap. 12). Note the large
instabilities generated, showing the poor performance of the scheme for resolving
rarefaction and shock waves in free surface flows. The small value of CFL used was
needed to obtain a computational result. For larger values, computations failed.
Attempts to use this scheme are found in the literature. A possible way to find an
“apparent” stable result is to introduce a large amount of dissipation in the scheme.
This is revealed in Fig. 7.9c, d, where the artificial viscosity method by Jameson
et al. (1981) was applied. Note that oscillations are suppressed, but this is no
indication that the scheme correctly works; simply, a numerical artifact was used to
force a solution.
Simulations conducted for the same test using the Lax diffusive scheme and
CFL ¼ 1 are presented in Fig. 7.10a, b. As expected, the shock and rarefaction
waves are smeared due to the diffusion term introduced by this scheme. The
numerical model is implemented in the code available on the file
“Dambreak_Diffusive.xls” (Chap. 12). The same simulation was conducted using
CFL ¼ 0:2, with the results shown in Fig. 7.10c, d. The low CFL value introduced
an unacceptable amount of numerical diffusion.
The simulation results from the MacCormack scheme are shown in Fig. 7.11a, b.
Spurious oscillations in the shock front and at the toe of the rarefaction wave are
present, albeit more visible in the velocity profile. These oscillations result from
dispersive errors. Applying the artificial viscosity method of Jameson et al. (1981)
as presented by Chaudhry (2008) with a calibrated K = 2, the oscillations are
almost suppressed, as noted from Fig. 7.11c, d. The numerical model is imple-
mented in the code available from “Dambreak_MacCormack_Jameson.xls”
(Chap. 12). Applying the MacCormack TVD scheme results in Fig. 7.12a, b.
7.5 Computational Examples 301
Fig. 7.9 Subcritical dam break wave for hd/hu = 0.4 computed with the unstable scheme using
CFL ¼ 0:1 showing effect of artificial viscosity coefficient K
Fig. 7.10 Subcritical dam break wave for hd/hu = 0.4 computed with diffusive scheme
highlighting effect of CFL
302 7 Finite Difference Methods
Fig. 7.11 Subcritical dam break wave for hd/hu = 0.4 computed with MacCormack scheme using
CFL ¼ 0:9 showing effect of artificial viscosity coefficient K
Fig. 7.12 Subcritical dam break wave for hd/hu = 0.4 computed with TVD MacCormack scheme
using CFL ¼ 0:9
Spurious oscillations in the shock front and at the toe of the rarefaction wave are
suppressed without the need of using any calibration parameter, which is a
remarkable advantage. The numerical model is implemented in the code available
on the file “Dambreak_MacCormack_TVD.xls” (Chap. 12). The simulation results
obtained using the upwind scheme are shown in Fig. 7.13a, b. Note that the dif-
fusion of the scheme, as in other first-order schemes, is present. The numerical
model is implemented in the code available on the file “Dambreak_Upwind.xls”
(Chap. 12).
7.5 Computational Examples 303
Fig. 7.13 Subcritical dam break wave for hd/hu = 0.4 computed with upwind scheme using
CFL ¼ 0:9
The subcritical dam break wave is not a severe test case (Zoppou and Roberts
2003). A transcritical dam break wave was thus generated using a depth ratio
hd/hu = 0.001 and hu = 1 m. Critical flow establishes at the dam axis in a point of
the rarefaction wave. Only the TVD MacCormack, upwind, and Lax diffusive
schemes were able to generate computational results, which are displayed in
Fig. 7.14 for t = 2 s.
The results of the MacCormack TVD scheme are excellent, as shown in
Fig. 7.14a, b. The upwind method, however, produces an unphysical jump at the
dam axis related to a violation of the entropy condition. The Lax diffusive scheme
produces results not in precise agreement neither for the rarefaction nor for the
shock waves.
U1 h1 U2 h2
Vw ¼ ; ð7:103Þ
h 1 h2
304 7 Finite Difference Methods
Fig. 7.14 Transcritical dam break wave for hd/hu = 0.001 computed using CFL ¼ 0:9 with
(a, b) MacCormack TVD scheme, (c, d) upwind scheme, (e, f) Lax diffusive scheme
Combining Eqs. (7.103) and (7.104), and using the definition of unit discharge
q = Uh, results in
1=2
q1 q2 g h1
¼ U2 þ ð h1 þ h2 Þ ; ð7:105Þ
h1 h2 2 h2
7.5 Computational Examples 305
or
1=2
g h1
q1 ¼ U2 h1 þ ðh1 h2 Þ ð h1 þ h2 Þ : ð7:106Þ
2 h2
This is an implicit equation to be solved for h1, once the values of q1, h2, and U2 are
prescribed, e.g.,
1=2
g h1
f ð h1 Þ U 2 h1 þ ð h1 h2 Þ ð h1 þ h2 Þ q1 ¼ 0: ð7:107Þ
2 h2
fz
hz1 þ 1 ¼ hz1 : ð7:108Þ
ðdf /dh1 Þz
Once h1 is determined, Vw follows from Eqs. (7.103) or (7.104), and the profile
given by Eq. (7.102) is determined. The solution is implemented in a code available
on the file “SurgeAnalytical.xls” (Chap. 12).
The MacCormack TVD scheme was used to simulate a subcritical surge prop-
agating over initially still water of depth 0.205 m. A sudden increase of discharge at
the inflow section of 0.028 m3/s was supplied in a channel 0.42 m wide. These are
the test conditions for surge experiments conducted by Favre (1935). Consider now
ideal surges (n = 0 m1/3/s). At the first time step, the water depth at the upstream
section was determined with the boundary conditions for discontinuous flows, to
apply then the method of characteristics for the remaining time steps with
Eq. (7.37) as
306 7 Finite Difference Methods
g kþ1 k
U1k þ 1 ¼ U2k þ k
h1 hk2 þ g So Sf 2 Dt: ð7:111Þ
c2
g k þ 1 2 g k k
h þ U k
h þ g S o S f 2 Dt hk1 þ 1 qk1 þ 1 ¼ 0: ð7:112Þ
ck2 1 2
ck2 2
the analytical solution except for a small hump numerically generated during the
initial simulation instants. To evidence the importance of correct implementation of
boundary conditions, we demonstrate the result of implementing naive approxi-
mations. It is tempting to assume that a boundary section is just another compu-
tational point, treating @F=@x using either backward of forward finite-differences. In
our case, the idea may be to generate the unknown water depth at the inflow section
by discretizing the continuity equation using forward finite-differences. That is,
Dt k
hk1 þ 1 ¼ hk1 q2 qk1 : ð7:113Þ
Dx
This naive determination of the unknown water depth at the inflow section was
implemented, and the poor results generated are shown in Fig. 7.16c, d, as found by
García-Navarro and Saviron (1992). This demonstrates the importance of the cor-
rect implementation of boundary conditions. The numerical solution is imple-
mented in the code “Favre_FDM_TVD_ BoundaryDiscont_Super_Cont.xls”
(Chap. 12).
Favre (1935) measured a subcritical surge propagating over initially still water of
depth 0.205 m. The discharge increase at the inlet section was 0.028 m3/s, and
the channel width was 0.42 m. Manning’s coefficient for this flume was
n = 0.01 m1/3/s (Favre 1935; Terzidis and Strelkoff 1970). Computational results
generated with the TVD MacCormack scheme along with a discontinuous–con-
tinuous implementation of the upstream boundary condition were generated at
several instants of time. Results are displayed in Fig. 7.17 and compared with
experimental data. The time needed to switch on the pumps and generate into the
flume the discharge pulse was computationally accounted for as a time lag of 1 s,
given that the generation of the discharge pulse is instantaneous in the mathematical
model. Comparing profiles at two different instants of time, the absolute velocity of
the surge predicted by the numerical model is in excellent agreement with that
indicated by the experiments. The numerical solution is implemented in the code
“Favre_FDM_TVD_ BoundaryDiscont_Subcrit.xls” (Chap. 12).
7.5 Computational Examples 309
References
Steger, J. L., & Warming, R. F. (1981). Flux vector splitting of the inviscid gas dynamic equations
with application to finite-difference methods. Journal of Computational Physics, 40(2),
263–293.
Stoker, J. J. (1957). Water waves: The mathematical theory with applications. New York:
Interscience Publishers.
Sturm, T. W. (2001). Open channel hydraulics. New York: McGraw-Hill.
Terzidis, G., & Strelkoff, T. (1970). Computation of open channel surges and shocks. Journal of
the Hydraulics Division, ASCE, 96(HY12), 2581–2610.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. Singapore: Wiley.
Toro, E. F. (2009). Riemann solvers and numerical methods for fluid dynamics. London: Springer.
Zoppou, C., & Roberts, S. (2003). Explicit schemes for dam-break simulations. Journal of
Hydraulic Engineering, 129(1), 11–34.
Chapter 8
The Riemann Problem
8.1 Introduction
@U @F
þ ¼ 0. ð8:2Þ
@t @x
Fig. 8.1 Initial conditions of Riemann problem for the SWE a physical plane, b x-t plane
Here, U is the vector of unknowns and F is the flux vector given, with h as the
water depth and U as the depth-averaged velocity, by
h Uh
U¼ ; F¼ : ð8:3Þ
Uh U 2 h þ 12 gh2
The fundamental difference with the dam break problem stems from the possi-
bility of having nonvanishing initial values of U at each side of the discontinuity.
The basic wave solutions are those described in Chaps. 5 and 6, namely shock
waves and rarefaction waves. Previous relations are then first generalized to
accommodate arbitrary initial data, and then, the solution strategy of the wave
structure for wet-bed conditions is detailed. The computation of the complete wave
profile is presented. The dry-bed Riemann solver is subsequently explained, and
finally, the instantaneous gate operations in open channels are analyzed in detail
using the theory presented.
8.2 Wet-Bed Exact Riemann Solver 315
As previously discussed, the basic wave solutions of Eqs. (8.1), (8.2) are shocks
and rarefaction waves. The type of wave generated at each side of the discontinuity
for t > 0 is not known in advance, demanding for an iterative solution of the wave
field starting with an assumed wave pattern. In general, four cases may arise:
(1) Left wave is a rarefaction, and right wave is a shock.
(2) Left wave is a rarefaction, and right wave is a rarefaction.
(3) Left wave is a shock, and right wave is a shock.
(4) Left wave is a shock, and right wave is a rarefaction.
Consider case 1, with a left-going rarefaction wave and a right-going shock
(Fig. 8.2). The basic equation for a left rarefaction wave results from the relation
stating conservation of the Riemann invariant U + 2c within the wave, that is, with
c = (gh)1/2 [Eq. (5.78)]
or,
h i
U ¼ UL 2ðc cL Þ ¼ UL 2 ðgh Þ1=2 ðghL Þ1=2 : ð8:6Þ
This equation states the conditions at the star region as function of initial conditions
at the left side of the discontinuity. It is therefore an equation “connecting” the
constant state zone with the left-side initial conditions. Following Toro (2001), we
define a left-side function fL as
U ¼ UL fL ;
h i ð8:7Þ
fL ¼ 2 ðgh Þ1=2 ðghL Þ1=2 :
The propagation speed SR of a right-going shock wave is given from the Rankine–
Hugoniot jump conditions by [Eq. (5.35)]
1=2
gh
SR ¼ UR þ ð h þ hR Þ : ð8:8Þ
2hR
hR ðUR SR Þ ¼ h ðU SR Þ; ð8:9Þ
it is possible to write
U ¼ UR þ fR ;
1=2 ð8:12Þ
g
f R ¼ ð h hR Þ ð h þ hR Þ :
2h hR
8.2 Wet-Bed Exact Riemann Solver 317
Here, fR is a function “connecting” the constant state zone with the right-side
conditions. Equating the velocity in the star region from Eqs. (8.7) and (8.12)
produces
This is an equation that must be numerically solved to determine the water depth at
the star region h*. Once done, U* is computed either from Eqs. (8.7) or (8.12). The
important aspect at this stage is that the Riemann problem produces an equation
where the constant state zone is linked to initial data via left- and right-hand
functions. Note that the type of function is different for shock and rarefaction waves,
thereby indicating that it is necessary to discriminate in advance which type of wave
originates at each side of the discontinuity. Note further that the left-going rar-
efaction wave produces a depression of water depths (h* < hL), whereas the
right-going shock involve an increase in water level (h* > hR). An important issue
is that Eq. (8.13) is general, as verified by repeating this analysis for the other three
types of wave patterns. It implies formulating the equations for a left-going shock
using the Rankine–Hugoniot jump conditions, and the continuous solution of the
right-going smooth wave, basically implying conservation of the Riemann invariant
U(x, t) − 2c(x, t) across the rarefaction (Jain 2001; Toro 2001). The repetition
process is elementary and is left to the reader. The results for fR and fL, depending
on whether they are linked to rarefactions of shocks, are summarized as (Toro 2001)
8 h i
< 2 ðgh Þ1=2 ðghL Þ1=2 if h hL ðrarefaction waveÞ
fL ¼ h i ; ð8:14Þ
: ðh h Þ g ðh þ h Þ 1=2 if h [ hL ðshock waveÞ
L 2h hL L
8 h i
< 2 ðgh Þ1=2 ðghR Þ1=2 if h hR ðrarefaction waveÞ
fR ¼ h i : ð8:15Þ
: ðh h Þ g ðh þ h Þ 1=2 if h [ hR ðshock waveÞ
R 2h hR R
Note that in a given Riemann problem, once the value of h* is adopted, the types of
waves generated at the right and left sides are automatically determined.
318 8 The Riemann Problem
The solution process is iterative, starting with an initial guess of h*. This value,
upon comparison with hL and hR, determines which types of waves generate at the
8.2 Wet-Bed Exact Riemann Solver 319
left- and right-hand sides. In general, the initial guess ho does not satisfy Eq. (8.13),
producing
UR UL þ fR ho ; hR þ fL ho ; hL 6¼ 0: ð8:16Þ
fk
ðh Þk þ 1 ¼ ðh Þk : ð8:17Þ
ðdf =dh Þk
With this new value of h*, the types of waves at L or R may change, therefore. Thus,
the wave field is iteratively updated in the solution process. The derivative term
needed in Eq. (8.17) is
df dfL dfR
¼ þ ; ð8:18Þ
dh dh dh
where
8
dfL < hgðgh Þ
1=2
if h hL ;
¼ i1=2 h i1=2 ð8:19Þ
ðh hL Þ
dh : 2hg h ðh þ hL Þ g 4h 2
g
2h hL ð h
þ h L Þ if h [ hL ;
L
and
8
dfR < gh ðgh Þ
1=2
if h hR ;
¼ i1=2 h i1=2 ð8:20Þ
ðh hR Þ
dh : 2hg h ðh þ hR Þ g 4h 2 2h
g
h ð h þ hR Þ if h [ hR :
R R
The wave relation across the left rarefaction implies conservation of the Riemann
invariant U + 2c, resulting in
320 8 The Riemann Problem
1
U ¼ ðUR þ UL Þ þ ðcL cR Þ: ð8:25Þ
2
SL ¼ UL kL ðghL Þ1=2 ;
h i
1=2 ð8:26Þ
kL ¼ 12 ðh þh2hL Þh :
L
This is needed to track the position of the shock front at any instant of time. The
depth and velocity within the shock wave are simply h* and U*.
Left rarefaction wave
For a left rarefaction wave, the signal speeds of its edge and tail are, respectively,
SL ¼ UL ðghL Þ1=2 ;
ð8:27Þ
SL ¼ U ðgh Þ1=2 :
Between the edge and the tail of the rarefaction, the spatial variations of depth and
velocity for any instant of time are given by solving x/t = U–c and U + 2c = const.
for U(x, t) and c(x, t) by
1h xi
ðghÞ1=2 ¼ UL þ 2ðghL Þ1=2 ;
3 t ð8:28Þ
1h 1=2 xi
U ¼ UL þ 2ðghL Þ þ 2 :
3 t
8.2 Wet-Bed Exact Riemann Solver 321
For a right shock wave, the signal speed is, from the Rankine–Hugoniot jump
conditions
SR ¼ UR þ kR ðghR Þ1=2 ;
ð8:29Þ
1 ðh þ hR Þh 1=2
kR ¼ ;
2 h2R
which is similarly needed to track the position of the shock front at any instant of
time. The depth and velocity within the shock wave are simply h* and U*.
Right rarefaction wave
For a right rarefaction wave, the signal speeds of its edge and tail are, respectively,
SR ¼ UR þ ðghR Þ1=2 ;
ð8:30Þ
SR ¼ U þ ðgh Þ1=2 :
Between the rarefaction edge and tail, the spatial variations of depth and velocity
for any instant of time are given by
1h xi
ðghÞ1=2 ¼ UR þ 2ðghR Þ1=2 þ ;
3 t ð8:31Þ
1h 1=2 xi
U ¼ UR 2ðghR Þ þ 2 :
3 t
For the four cases previously reported, there are a number of subcases depending on
the relative position of each wave with respect to the t-axis, determined by the signs
of SL, SL , SR and SR . The 16 possible cases are plotted in Figs. 8.3, 8.4, 8.5, and 8.6
(Zoppou and Roberts 2003; Roberts 2013).
As stated at the start of this chapter, the Riemann problem is a key ingredient
used in Godunov-type finite volume methods. Here, it is advanced that a funda-
mental quantity to be determined is the value of the flux vector F at coordinate
x = 0, i.e., the flux crossing the plane where the discontinuity in U was initially
generated. Using the definitions of U and F, it is elementary to demonstrate that
322 8 The Riemann Problem
Fig. 8.3 Wave patterns of Riemann problem for left rarefaction wave and right shock wave
(case 1) (adapted from Roberts 2013)
h U1 U2
U¼ ¼ ; F ¼ FðUÞ ¼ U22 : ð8:32Þ
Uh U2 U1 þ 12 gU12
This means that the value of F is fully determined by U. Thus, the value of F(x = 0)
is given by the solution of the Riemann problem at x = 0, namely U(x = 0).
Depending on the specific wave pattern, U(x = 0) may equal UL, UR, U*, or a value
Uc to be determined using the rarefaction wave equations, corresponding to critical
flow. The computation of the value of F(x = 0) for the 16 possible cases depicted in
Figs. 8.3, 8.4, 8.5 and 8.6 is summarized in Table 8.1. Note that some cases are
equivalent in terms of computation of F(x = 0), despite the different complete wave
patterns. If a finite volume scheme is constructed using the exact solution of the
Riemann problem, it is said to use an “exact Riemann solver.” This nomenclature
should be understood as the use of the numerical solution originating from the
consideration of the complete Riemann problem. “Exact Riemann solver” should
not be confused with having closed-form analytical solutions in a Riemann solver.
In fact, this is only true for the two-rarefaction case. To the authors’ knowledge,
exact analytical solutions are unknown for the rest of wave patterns.
8.2 Wet-Bed Exact Riemann Solver 323
Fig. 8.4 Wave patterns of Riemann problem for left and right rarefaction waves (case 2) (adapted
from Roberts 2013)
Fig. 8.5 Wave patterns of Riemann problem for left and right shock waves (case 3) (adapted from
Roberts 2013)
Fig. 8.6 Wave patterns of Riemann problem for left shock wave and right rarefaction wave
(case 4) (adapted from Roberts 2013)
8.2 Wet-Bed Exact Riemann Solver 325
Table 8.1 Evaluation of the flux F in the Riemann problem (see Figs. 8.3, 8.4, 8.5, and 8.6)
Case Subcase F(x = 0)
1 a F(Uc); critical flow at left rarefaction
1 b F(U*)
1 c F(UL)
1 d F(UR)
2 a F(Uc); critical flow at right rarefaction
2 b F(U*)
2 c F(UL)
2 d F(UR)
2 e F(Uc); critical flow at left rarefaction
3 a F(U*)
3 b F(UR)
3 c F(UL)
4 a F(Uc); critical flow at right rarefaction
4 b F(UR)
4 c F(UL)
4 d F(U*)
Table 8.2 Test conditions for examples of wave solutions of the Riemann problem
Test Time (s) hL (m) hR (m) UR (m/s) UL (m/s)
Figure 8.7 2 0.2 0.4 0 7
Figure 8.8 2 1 1 5 –5
Figure 8.9 2 1 0.2 2.5 0.5
Fig. 8.7 Riemann solution of test with two right-going shocks for a free surface profile h(x) and
b depth-averaged velocity U(x)
Table 8.3 Test results for examples of wave solutions of Riemann problem. The discontinuity is
located at x = 0, t = 0
Test Time (s) SR (m/s) SL (m/s) h* (m) U* (m/s)
Figure 8.7 2 4.213 1.477 1.02 2.56
Figure 8.8 2 8.132 −8.132 0.041 0
Figure 8.9 2 4.5 −2.632 0.317 3.238
326 8 The Riemann Problem
Fig. 8.8 Riemann solution of test with left- and right-going rarefaction waves for a free surface
profile h(x) and b depth-averaged velocity U(x)
directions. The interest of this type of test is that the interaction of the rarefaction
waves produces a star region of static water where the bed may become dry under
certain conditions. The mathematical conditions for two identical rarefaction waves
traveling in opposite directions are hR = hL and UL = −UR. Inserting these into
Eq. (8.22) produces
" #2 2
1 2ðghR Þ1=2 1 1 1
h ¼ ð2UR Þ ¼ ðghR Þ1=2 UR : ð8:33Þ
g 2 4 g 2
1
U ¼ ðUR UR Þ þ ðcR cR Þ ¼ 0: ð8:35Þ
2
Fig. 8.9 Riemann solution of test with left-going rarefaction wave and right-going shock for
a free surface profile h(x) and b depth-averaged velocity U(x)
The wave structure of the Riemann problem changes drastically if a portion of the
computational domain is dry. Before presenting the corresponding wave solutions,
consider Toro (2001), who demonstrated that a shock wave cannot connect the wet–
dry portions. Let UL represent the initial wet-bed data, and UR the data of the
dry-bed region. Obviously, hR = 0. Assume that a shock wave propagating with
velocity S connects the two states. The Rankine–Hugoniot jump conditions are then
Setting hR = 0, the first identity yields S = UL. Substitution in the second identity
yields
hL UL2 þ 12 gh2L
S¼ UL ) hL ¼ 0; ð8:37Þ
hL U L
Consider first wet-bed conditions on the left, corresponding to the initial data
(Fig. 8.10)
328 8 The Riemann Problem
UL if x\0
Uðx; 0Þ ¼ : ð8:38Þ
0 if x[0
SR ¼ UR ¼ UL þ 2cL : ð8:41Þ
The depth and velocity within the rarefaction wave are thus given by (Toro 2001)
1h xi
ðghÞ1=2 ¼ UL þ 2ðghL Þ1=2 ;
3 t ð8:42Þ
1h 1=2 xi
U ¼ UL þ 2ðghL Þ þ 2 :
3 t
For wet-bed conditions on the right side (Fig. 8.11), the corresponding initial
data are
0 if x\0
Uðx; 0Þ ¼ : ð8:43Þ
UR if x[0
Fig. 8.12 Dry-bed conditions along center of two rarefactions a free surface profiles h(x) and
b x-t plane
The propagation velocity of the left-going wet–dry interface is determined using the
basic equation of the rarefaction wave, namely
SL ¼ UL ¼ UR 2cR : ð8:46Þ
The depth and velocity within the rarefaction wave are thus given by (Toro 2001)
1h xi
ðghÞ1=2 ¼ UR þ 2ðghR Þ1=2 þ ;
3 t ð8:47Þ
1h 1=2 xi
U ¼ UR 2ðghR Þ þ 2 :
3 t
Given that c* must be positive to generate real values of h*, the depth-positivity
condition is thus, for arbitrary data (Toro 2001),
1h i 1
c ¼ ðghL Þ1=2 þ ðghR Þ1=2 ðUR UL Þ 0: ð8:51Þ
2 4
Eleuterio Francisco Toro was born on July 16, 1946, at Capitan Pastene,
Chile. He started, in 1967, his career as Teacher of primary education at
Victoria, Chile, continuing studies of mathematical pedagogy, obtaining the
B.Sc. degree in pure mathematics in 1977 at the University of Warwick, UK,
and in 1978 the M.Sc. degree in applied mathematics from University of
Dundee, UK. He obtained the PhD degree in 1982 in computational mathe-
matics from the University of Teesside, Middlesbrough UK. From 1983 until
2001, he was Lecturer and Senior Lecturer in computational fluid mechanics
at Cranfield University, UK, and Professor of applied mathematics at the
Manchester Metropolitan University. From 2002 to 2016, he was Full
332 8 The Riemann Problem
Consider steady flow at a gate (Fig. 8.13), where subcritical approach flow (state L)
is transformed into supercritical flow1 (state R). Contraction effects are overlooked
in the present shallow water analysis (Jain 2001). These were accounted for by
Cozzolino et al. (2015) for both complete and partial gate openings.
If the gate is fully opened, the initial data correspond to the Riemann problem
stated in Fig. 8.1, with ULhL = URhR = const. as the initial discharge. Figure 8.2
shows the wave profile for a transcritical rarefaction after full gate opening. The
computational example presented in Fig. 8.9 represents thus a full gate opening.
1
We will analyze instantaneous (partial or full) sluice gate openings and closures for supercritical
conditions downstream of the gate. Submerged flows are not considered; the gate acts thus as a
control section in all cases.
8.4 Application: Gate Maneuvers in Open Channels 333
Fig. 8.13 Planar gate in open channel forming a discontinuity between two uniform flow zones.
Contraction effects are overlooked in the SWE analysis
Consider initially steady gate flow as sketched in Fig. 8.14a, with subscripts u and
R referring to the up- and downstream zones of the gate. For given values of hu and hR,
the values of Uu and UR are determined with the known initial discharge. For partial
gate opening (w > hR and w < hL) (Fig. 8.14b), a depression of depth hd and velocity
Ud is formed upstream of the gate, propagating in the upstream direction via a rar-
efaction wave. The unknowns in the upstream flow portion are thus hd, Ud and the
velocity at the gate opening UL. Across the rarefaction, one may write the identity
Two additional equations are needed. One involves energy conservation across the
gate
Ud2 U2
hd þ ¼ wþ L ; ð8:53Þ
2g 2g
UL w ¼ Ud hd : ð8:54Þ
UL2 U2 w2 U 2
¼ hd w þ d ¼ hd w þ 2 L ; ð8:56Þ
2g 2g hd 2g
334 8 The Riemann Problem
Fig. 8.14 Partial gate opening a initial steady state, b generation of transient motion (flow
downstream of the gate corresponds to case c of Fig. 8.3) (adapted from Montuori and Greco
1973)
or
1=2
1=2 w2
UL ¼ ½2gðhd wÞ 1 2 : ð8:57Þ
hd
Inserting Eq. (8.57) into Eq. (8.55) yields the nonlinear implicit equation for hd
Table 8.4 Example of partial gate opening. Initial discharge is 0.5 m2/s and new gate opening
w = 0.4 m
Flow hu (m) Uu (m/s) hd (m) Ud (m/s) UL (m/s)
Upstream 1 0.5 0.645 1.733 2.795
hR (m) UR (m/s) h* (m) U* (m/s)
Downstream 0.2 2.5 0.316 3.234
Fig. 8.15 Riemann solution of partial gate opening test for flow downstream of the gate plane
with a free surface profile h(x) and b depth-averaged velocity U(x). Details see Table 8.4
336 8 The Riemann Problem
Fig. 8.16 Partial gate opening a transient motion with two rarefactions beyond the gate,
corresponding to case c in Fig. 8.4, b transient with two shocks beyond the gate, corresponding to
case c in Fig. 8.5 (adapted from Montuori and Greco 1973)
Table 8.5 Example of partial gate opening. Initial discharge is 0.5 m2/s and new gate opening
w = 0.25 m
Flow hu (m) Uu (m/s) hd (m) Ud (m/s) UL (m/s)
Upstream 1 0.5 0.827 1.066 3.53
hR (m) UR (m/s) h* (m) U* (m/s)
Downstream 0.2 2.5 0.307 3.185
1=2
g
2ð h hR Þ ð h þ h R Þ ¼ UL UR : ð8:59Þ
2h hR
Note that the right-hand side of Eq. (8.59) is positive; to have a compatible
positive left-hand side, of necessity then h* > hR, implying that the two-shock
8.4 Application: Gate Maneuvers in Open Channels 337
Fig. 8.17 Riemann solution of partial gate opening test for flow downstream of the gate plane
with a free surface profile h(x) and b depth-averaged velocity U(x). Details see Table 8.5
assumed wave pattern, is the correct solution. Another simulation was conducted
based on the example presented in Table 8.4, but using the smaller gate opening
w = 0.25 m. Results are summarized in Table 8.5, with the flow profiles down-
stream of the gate presented in Fig. 8.17.
Consider the case of a complete gate closure (Fig. 8.18). Upstream of the gate, a
shock wave traveling to the left, is formed. The unknowns in this portion of the
computational domain are the velocity Ud, the flow depth hd, and the surge velocity
S (Jain 2001). By definition of gate closure, Ud = 0. Thus, the two remaining
unknowns are determined using the mass and momentum conservation equations
across the surge, e.g., the Rankine–Hugoniot jump conditions, which are,
respectively,
Fig. 8.18 Complete gate closure a wet bed beyond the gate, b dry bed beyond the gate
meaningless. The simulations were conducted with the wet-bed Riemann solver
available on the file “ExactRiemannSolver_wetbed.xls”. A second test was conducted
by increasing the velocity at the tailwater section to produce a Froude number in excess
of 2. A dry-bed two-rarefaction approach was used to model the flow, with the profiles
for x > 0 plotted in Fig. 8.19c, d. The results were generated with a dry-bed Riemann
solver available on the file “ExactRiemannSolver_drybed.xls”.
For the case of a partial gate closure, a surge propagates in the upstream direction
within the approach flow to the gate (Fig. 8.20). The unknowns in this flow portion
are S, hd, Ud, and UL. The equations available are the continuity equation across the
surge
8.4 Application: Gate Maneuvers in Open Channels 339
Fig. 8.19 Riemann solution for full gate closure test: flow downstream of the gate with a, b free
surface profile h(x) and depth-averaged velocity U(x) for wet-bed conditions, c, d idem for dry-bed
conditions. Details see Table 8.6
Table 8.6 Examples of complete gate closure using the two-rarefaction Riemann problem
Test hL (m) hR (m) UL (m/s) UR (m/s) h* (m) U* (m/s)
Wet-bed 1.2 1.2 −5 5 0.088 0
Dry-bed 1.2 1.2 −10 10 0 3.138
Ud2 U2
hd þ ¼ wþ L ; ð8:64Þ
2g 2g
UL w ¼ Ud hd : ð8:65Þ
340 8 The Riemann Problem
Fig. 8.20 Partial gate closure a shock–rarefaction waves, see case c in Fig. 8.6, b two shock
waves, see case c in Fig. 8.5, c two rarefaction waves, see case c in Fig. 8.4 (adapted from
Montuori and Greco 1973)
8.4 Application: Gate Maneuvers in Open Channels 341
Once this system of four equations and four unknowns is solved (Jain 2001), the
flow downstream of the gate is computed as a Riemann problem. The case
involving a shock wave followed by a rarefaction wave is plotted in Fig. 8.20a.
Depending on the initial data of the Riemann problem, a number of wave profiles
may be generated beyond the gate. The cases of two shocks and two rarefaction
waves are sketched in Fig. 8.20b, c, respectively.
These scenarios are illustrated with a numerical example using the code avail-
able in “ExactRiemannSolver_wetbed.xls”. Data for this example are taken from
Jain (2001, example 7-6). The initial discharge is qo = 7.92 m2/s, hR = 1 m,
UR = qo/hR = 7.92 m/s and the gate opening is set to w = 0.7 m. Solving the
equations upstream of the gate yields UL = 8.63 m/s (Jain 2001). The downstream
flow profiles for this test are plotted in Fig. 8.21a, b, and computations are sum-
marized in Table 8.7. Note that the wave patterns involve a left shock and right
rarefaction. Henceforth, we keep w fixed. Given that the gate acts as a control, flow
conditions upstream of the gate remain unaltered for fixed w. Now, raise hR up to
1.2 m, computing the corresponding UR to satisfy the initial data. As observed from
Figs. 8.21c, d and Table 8.7, the wave patterns in this case correspond to two
shocks. If the analysis is repeated this time lowering hR to 0.8 m, two rarefactions
waves travel to the right of the gate (Fig. 8.21e, f and Table 8.7). As illustrated in
this example, the treatment of the gate maneuvers within the context of a Riemann
problem permits to identify all possible wave patterns without the need of pre-
liminary assumptions.
The experimental investigation by Montuori and Greco (1973) is used to test the
overall quality of the theory presented against experiments. Two experiments are
selected, with experimental data for the Riemann problem posed beyond the gate
listed in Table 8.8, as well as visual observations reported with photographs by
Montuori and Greco (1973). Computations of the star region values using the
wet-bed Riemann solver are included in the same table. In the comparison under-
taken below, the left conditions of the Riemann problem are taken as the experi-
mental conditions at the new vena contracta after the gate movement. Thus,
contraction effects are accounted for here.
The first experiment is reproduced computationally in Fig. 8.22. Wave results
are plotted in Fig. 8.22 after 5 s of routing. Note that the experimental data for the
Riemann problem produce computationally a two-shock wave pattern, in concor-
dance with the visual observations by Montuori and Greco (1973). Table 8.9
compares the theory with available experimental measurements, indicating fair
agreement.
The second experiment is reproduced computationally in Fig. 8.23. Wave results
are plotted in Fig. 8.23 after 5 s of routing. The experimental data for the Riemann
problem in this test produce computationally a shock–rarefaction wave pattern, in
342 8 The Riemann Problem
Fig. 8.21 Partial gate closure. Generation of different wave patterns keeping w fixed and varying
hR (and UR = qo/hR). See Table 8.7 for test conditions of each plot
Table 8.7 Examples of Riemann problems downstream of gate after partial gate closure
Test Waves hR (m) UR (m/s) h* (m) U* (m/s)
Figure 8.21a, b Shock–rarefaction 1 7.92 0.95 7.76
Figure 8.21c, d Two shocks 1.2 6.6 1.26 6.77
Figure 8.21e, f Two rarefactions 0.8 9.9 0.58 9.08
Test conditions are: qo = 7.92 m2/s, hL = w = 0.7 m, UL = 8.63 m/s, qnew = ULw = 6.041 m2/s
Table 8.8 Experiments of Montuori and Greco (1973): data of Riemann problem beyond the gate
and computed values of star region using wet-bed Riemann solver
Test Wave pattern hL (m) hR UL UR h* U*
experimentally observed (m) (m/s) (m/s) (m) (m/s)
Figure 8.22 Two bores 0.057 0.1 1.844 0.11 0.165 0.685
Figure 8.23 A bore followed by smooth 0.0248 0.093 2.513 1.955 0.073 1.733
wave
Subscript L indicates conditions at the new vena contracta after gate movement, and R indicates
tailwater conditions existing before moving the gate
8.4 Application: Gate Maneuvers in Open Channels 343
Fig. 8.22 Computational reproduction of a two-shock wave pattern beyond a gate generated
experimentally by Montuori and Greco (1973). For details, see Table 8.9
Table 8.9 Comparison of theory and experiments by Montuori and Greco (1973) for the
Riemann problem presented in Fig. 8.22
SL (m/s) SR (m/s) h* (m)
Simulation 0.072 1.573 0.165
Experiment 0.1 1.45 0.165
Table 8.10 Comparison of theory and experiments by Montuori and Greco (1973) for the
Riemann problem presented in Fig. 8.23
SL (m/s) h* (m)
Simulation 1.329 0.073
Experiment 1.45 0.065
In this case, the p-characteristics are impinging on the discontinuity, while the other
characteristics are crossing it,
In our case, p can be 1 or 2, so one of the following set of identities shall be verified
by the depression shock:
Conditions for p = 2
k2 ðUL Þ [ S [ k2 ðUR Þ;
k1 ðUL Þ\S; ð8:69Þ
k1 ðUR Þ\S:
It can be easily verified that this set of statements is coherent with Fig. 7.3 and
Eq. (7.22).
Conditions for p = 1
k1 ðUL Þ [ S [ k1 ðUR Þ;
k2 ðUL Þ [ S; ð8:70Þ
k2 ðUR Þ [ S:
8.4 Application: Gate Maneuvers in Open Channels 345
Table 8.11 Verification of Lax entropy condition for depression shock in Fig. 8.23
(SL = 1.329 m/s), using data in Table 8.8
Test k1(UL) (m/s) k1(U*) (m) k2(UL) (m/s) k2(U*) (m/s)
Figure 8.23 2.02 0.889 3.007 2.578
This is the set of identities verified by the depression shock in Fig. 8.23. To analyze
this shock, the left state is that of the Riemann problem and the right state corre-
sponds to the star region. Therefore, we will check the identities
k1 ðUL Þ [ SL [ k1 ðU Þ;
k2 ð U L Þ [ SL ; : ð8:71Þ
k2 ð U Þ [ SL :
This work is accomplished in Table 8.11 using previous information of Table 8.8,
resulting that Eq. (8.71) is verified, and therefore, the depression shock in Fig. 8.23
is a correct physical solution, as also observed experimentally by Montuori and
Greco (1973).
The described exact solutions apply to analyze instantaneous gate operations. If
the gate maneuvers are not instantaneous, or if source terms are introduced into the
SWE to account, for example, for frictional effects, the theory described is invalid
and a numerical solution is then required.
References
Cozzolino, L., Cimorelli, L., Covelli, C., Della Morte, R., & Pianese, D. (2015). The analytic
solution of the shallow-water equations with partially open sluice-gates: The dam-break
problem. Advances in Water Resources, 80(6), 90–102.
Godunov, S. K. (1959). A difference method for numerical calculation of discontinuous solutions
of the equations of hydrodynamics. Matematicheskii Sbornik, 47(3), 271–306 (in Russian).
Guinot, V. (2003). Godunov-type schemes: An introduction for Engineers. Amsterdam, Boston:
Elsevier science.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan Co.
Hoffman, J. D. (2001). Numerical methods for engineers and scientists (2nd ed.). New York:
Marcel Dekker.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Jeppson, R. (2011). Open channel flow: Numerical methods and computer applications. CRC
Press, Taylor and Francis, New York.
Katopodes, N. D. (2019). Free surface flow: Computational methods. Oxford, UK:
Butterworth-Heinemann.
LeVeque, R. J. (2002). Finite volume methods for hyperbolic problems. New York: Cambridge
University Press.
Montuori, C. (1968). Brusca immissione di una corrente ipercritica a tergo di altra preesistente
[Sudden perturbation of a supercritical flow over the pre-existing flow]. L’Energia Elettrica,
45(3), 174–187 (in Italian).
346 8 The Riemann Problem
Montuori, C., & Greco, V. (1973). Fenomeno di moto vario a valle di una paratoia piana [Varied
flow phenomena beyond a plane gate]. L’Energia Elettrica, 50(2), 73–88 (in Italian).
Riemann, B. (1860). Über die Fortpflanzung ebener Luftwellen von endlicher Schwingungsweite
[On the propagation of plane air waves of finite amplitude]. Abhandlungen der Königlichen
Gesellschaft der Wissenschaften zu Göttingen, 8, 43–65 (in German).
Roberts, S. (2013). Numerical solution of conservation laws applied to the shallow water
equations. Lecture notes. Australia: Mathematical Sciences Institute, Australian National
University.
Stoker, J. J. (1957). Water waves: The mathematical theory with applications. New York:
Interscience publishers.
Toro, E. F. (1997). Riemann solvers and numerical methods for fluid dynamics. London: Springer.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. Singapore: Wiley.
Zoppou, C., & Roberts, S. (2003). Explicit schemes for dam-break simulations. Journal of
Hydraulic Engineering, 129(1), 11–34.
Chapter 9
Finite Volume Methods
9.1 Introduction
The SWE are solved in this chapter using finite volume methods for hyperbolic
conservation laws, where the space–time plane is divided into control volumes or
simply finite volumes. The presentation focuses on Godunov-type methods, in
which the solution of a number of local Riemann problems between each pair of
adjacent cells is used. The constructed numerical method satisfies locally the
conservation laws, so that wave propagation information is generated at the cell
interfaces. This wave information is used to construct the numerical scheme and
then evolve the solution in time, representing therefore an upwind scheme. First,
Godunov-type methods are stated for the homogeneous SWE, after which
approximate Riemann solvers are presented. The treatment of dry beds is discussed,
as well as the inclusion of source terms including the discretization of the bed slope
term to produce a well-balanced scheme. The treatment of bed friction is explained
in detail. The one-sided first-order method is presented, as well as the second-order
Total Variation Diminishing (TVD) MUSCL (Monotonic Upstream Centered
Schemes for Conservation Laws)-Hancock scheme. Numerical computations for
selected test cases are finally compared with exact solutions and experimental
observations to highlight the quality of computations.
@U @F
þ ¼ S: ð9:1Þ
@t @x
Here, U is the vector of the conserved variables, F the flux vector and S the source
term vector, given for a rectangular channel by
h hU 0
U¼ ; F¼ ; S¼ ; ð9:2Þ
hU hU 2 þ 12 gh2 gh @z
@x ghSf
b
with h as the water depth, U the depth-averaged velocity, zb the bed elevation, g the
gravity acceleration and Sf the friction slope. The differential Eq. (9.1) is valid in
zones of the computational domain with smooth or continuous solutions, but it does
not apply at discontinuous portions as shocks. Therefore, Eq. (9.1) is integrated
over a control volume in the x-t plane, resulting in
ZZ ZZ
@U @F
þ dxdt ¼ Sdxdt: ð9:3Þ
@t @x
The integral Eq. (9.3) allows for the computation of both continuous and discontinuous
solutions. It is the fundamental relation of the finite volume method. For the rectangular
control volume in the x-t plane depicted in Fig. 9.1, one can write (Toro 2001, 2009)
tZþ Dt xZ
i þ 1=2 Z
xi þ 1=2 tZþ Dt tZþ Dt Z
xi þ 1=2
Z Z
xi þ 1=2 t þ Dt
@U @F @U @F
þ dxdt ¼ dx dt + dt dx ¼ Sdxdt:
@t @x @t @x
t xi1=2 xi1=2 t t xi1=2 xi1=2 t
ð9:4Þ
Here, i is the cell index in the x-direction, and i + 1/2 the interface between cells
i and i + 1. Equation (9.4) generates after elementary integration (Toro 2001, 2009)
Z
xi þ 1=2 tZþ Dt
½Uðx; t þ DtÞ Uðx; tÞdxþ F xi þ 1=2 ; t F xi1=2 ; t dt
xi1=2 t
ð9:5Þ
Z Z
xi þ 1=2 t þ Dt
¼ Sdxdt:
xi1=2 t
tZþ Dt
1
Fi þ 1=2 ¼ F xi þ 1=2 ; t dt;
Dt
t
tZþ Dt
1
Fi1=2 ¼ F xi1=2 ; t dt;
Dt
t
xi1=2 þ Dx
Z
1
Uki þ 1 ¼ Uðx; t þ DtÞdx; ð9:6Þ
Dx
xi1=2
xi1=2 þ Dx
Z
1
Uki ¼ Uðx; tÞdx;
Dx
xi1=2
Z Z
xi þ 1=2 t þ Dt
1
Si ¼ Sdxdt:
DxDt
xi1=2 t
Inserting Eq. (9.6) into Eq. (9.5) results in the exact conservative formula
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 þ DtSi : ð9:7Þ
Dx
This is the fundamental equation of the schemes to be presented here. The com-
putational domain in the finite volume method is divided into a number of control
350 9 Finite Volume Methods
volumes in the x-t plane (Fig. 9.2), where Eq. (9.7) is applied. Here, Fi+1/2 is the
numerical flux across the interface between cells i and i + 1. No numerical
approximations are introduced in Eq. (9.7), although it looks like a finite-difference
equation. Until stating the contrary, the homogeneous version of the SWE is
focused by dropping the source terms, e.g., setting Si = 0 in Eq. (9.7).
Note that in the conservative formula (9.7), the flux leaving cell i through the
interface i + 1/2 is identical to the flux entering into cell i + 1 across this common
interface. Let us write the conservative formula for updating the conserved variables
at cell i − 1,
þ1 Dt
Uki1 ¼ Uki1 Fi1=2 Fi3=2 : ð9:8Þ
Dx
Dt
Uki þ 1 þ Uki1
þ1
¼ Uki þ Uki1 Fi þ 1=2 Fi3=2 : ð9:9Þ
Dx
It indicates that the total variation of U is determined by the fluxes entering and
leaving the extreme interfaces of the block formed by the adjacent cells i − 1 and
i. If the computational domain is divided into N cells, the summation of the
updating equation for all the cells yields
9.2 Godunov-Type Schemes 351
X
N X
N
Dt
Uki þ 1 ¼ Uki FN þ 1=2 F1=2 : ð9:10Þ
i¼1 i¼1
Dx
tZþ Dt
1
qi þ 1=2 ¼ Fi þ 1=2 ð1Þ ¼ q xi þ 1=2 ; t dt; ð9:11Þ
Dt
t
corresponding to a time average of the discharge crossing the interface i + 1/2. This
definition of discharge satisfies the conservative property, for the reasons stated
above. However, the numerical discharge can also be defined as the second com-
ponent of the vector U; it is a spatial average value at the time level k + 1,
xi1=2 þ Dx
Z
1
qki þ 1 ¼ Uki þ 1 ð2Þ ¼ qðx; t þ DtÞdx: ð9:12Þ
Dx
xi1=2
This definition of the numerical discharge does not satisfy the conservative prop-
erty, however. Whether Eq. (9.11) or (9.12) for the numerical discharge is used to
represent the true physical discharge is open to debate. In some steady flow
problems (q = const.), like the hydraulic jump, the steady-state solution generated
by the SWE using certain numerical schemes for transient computations experience
unphysical jumps in qki þ 1 , whereas qi þ 1=2 attains a constant steady value due to the
conservative property (Ying and Wang 2008). This last definition of numerical
discharge may be then more useful to approximate the true physical discharge
(Ying and Wang 2008). In practice, computational schemes are not affected by the
choice; it is only a matter on which variable is taken as the output numerical
discharge once the solution is determined.
The conservative Eq. (9.7) with Si = 0 applies to update the solution of the
space-averaged values Ui once the numerical flux Fi+1/2 is estimated. This step
transforms the exact Eq. (9.7) approximate for a particular computation. The exact
solution U(x, t) within a cell can be written in the form of a Taylor series as
352 9 Finite Volume Methods
@U ðx xi Þ2 @ 2 U
UðxÞ ¼ Uðxi Þ þ ðx xi Þ þ þ :::: ð9:13Þ
@x i 2 @x2 i
In the Godunov upwind method, the actual solution U(x) at time t is approximated
by the space-averaged values within each cell (Fig. 9.3)
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 : ð9:15Þ
Dx
Consider two initial states UL and UR at a generic interface i + 1/2 between cells
i and i + 1. This is by definition a local Riemann problem, from which a number of
wave patterns may result. The typical case with a left-going rarefaction wave and
right-going shock is presented in Fig. 9.5. The shock and rarefaction waves
propagate with signal speeds SR and SL, respectively. The exact Riemann solution
presented in Chap. 8 thus applies here, taking the role of a component of the
numerical scheme. The complete Riemann solution is constructed assembling rar-
efaction and shock waves. Thus, it is noted that the solution of the Riemann
problem is self-similar, e.g., Ui+1/2(x/t). At the original position of the discontinuity
(x = 0), the solution of the Riemann problem is thus independent of t and therefore
steady. The constant state region behind the shock front (star region) depicted in
Fig. 9.5 is denoted as U*. The numerical flux crossing the original discontinuity at
x = xi+1/2 is Fi+1/2, which is needed to apply Eq. (9.15). The star region in the
Riemann problem is a steady-state zone where the conserved variables are U*. This
is part of the total (local) Riemann solution Ui+1/2. Therefore, the numerical flux
crossing the t-axis in the Riemann problem is also a constant, so that the intercell
flux is exactly evaluated based on the Riemann solution at x = xi+1/2 as
Ztk þ 1 ZDt
1 1
Fi þ 1=2 ¼ F x ¼ xi þ 1=2 ; t dt ¼ Fðx ¼ 0; tÞdt
Dt Dt
tk
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
0
|fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl} ð9:16Þ
global system of reference local system of reference
¼ Fðx ¼ 0Þ ¼ F Ui þ 1=2 ðx ¼ 0Þ :
This is the Godunov upwind numerical flux, namely the physical flux function
evaluated using the solution of the Riemann problem at the location of the interface
in question. Note that in local coordinates, the origin of the discontinuity is taken at
x = 0.
For the case depicted in Fig. 9.5, the vector Ui+1/2(x = 0) is U*. Thus,
The value of U* is determined solving the exact Riemann problem for wet-bed
conditions. If the left rarefaction wave crosses the t-axis, critical flow occurs along
it, producing the values of c and U after setting dx/dt = U − c = 0 (Toro 2001),
1
cc ¼ ðghc Þ1=2 ¼ Uc ¼ ðUL þ 2cL Þ: ð9:18Þ
3
1
hc ¼ ðUL þ 2cL Þ2 : ð9:19Þ
9g
This is also a steady vector. For a right-going rarefaction the flow is simply neg-
ative. Thus, critical flow is obtained setting dx/dt = U + c = 0, resulting in (Toro
1992, 2001)
1
cc ¼ ðghc Þ1=2 ¼ Uc ¼ ðUR þ 2cR Þ; ð9:21Þ
3
1
hc ¼ ðUR þ 2cR Þ2 : ð9:22Þ
9g
which is obviously identical to Eq. (9.20) but with a flow reversal. If the left-state
values reach the t-axis, the numerical flux is
9.2 Godunov-Type Schemes 355
hL UL
Fi þ 1=2 ¼ FðUL Þ ¼ : ð9:24Þ
hL UL2 þ 12 gh2L
Consider the right wave from the Riemann problem formed at interface i − 1/2 and
the left wave from Riemann problem formed at interface i + 1/2, both propagating
within cell i (Fig. 9.6). Consider that the wave of maximum speed Smax within the
data is the left wave emanating from interface i + 1/2. For stability of the explicit
Godunov scheme, the Courant–Friedrichs–Lewy number CFL must be limited. The
time step Dt is determined at time level k using the equation
Dx
Dt ¼ CFL: ð9:26Þ
Smax
1 Dx
Dt ¼ : ð9:27Þ
2 Smax
This time step avoids that during time Δt, the fastest signal crosses more than half of
the cell width (see point A in Fig. 9.6a). However, a careful inspection of the
conservative Eq. (9.15) and the Godunov numerical flux Eq. (9.16) reveal that
computations are only a function of the numerical flux Fi+1/2, which, in turn, is not a
356 9 Finite Volume Methods
function of the entire Riemann solution, but rather of the particular value at xi+1/2.
Thus, the detailed Riemann solution is not important at all, and wave interaction
within a cell can be permitted as long as the waves originating from a discontinuity
are not affecting the solution of the Riemann problem at the adjacent discontinu-
ities. The stability condition of the scheme is thus CFL ¼ 1, providing a maximum
time step of (Fig. 9.6b, point B)
Dx
Dt ¼ : ð9:28Þ
Smax
In practice, CFL ¼ 0:9 is usually adopted. One may estimate Smax using the
eigenvalues available at time k, e.g.,
9.2 Godunov-Type Schemes 357
Dx Dx
Dt ¼ CFL ¼ h i CFL: ð9:29Þ
Smax 1=2
max jU j þ ðghk Þ
k
i i
Under the presence of dry fronts, whose speeds are much faster than the eigen-
values, the scheme may turn unstable, so care is required. In these cases, a lower
value of CFL may be necessary to produce stable results.
2. Solve exactly the Riemann problem R(Ui, Ui+1) formed at interface i + 1/2
between cells i and i + 1. The complete solution Ui+1/2(x/t) of each Riemann
problem will be used. The numerical flux is not determined. Repeat for all
computational cells
3. Use Eq. (9.29) to determine a stable time step Δt, with CFL\1=2, given that
the complete solution of the Riemann problem will be used. This limit avoids
that the waves emanating from an interface interact with waves originated at
adjacent interfaces. Interaction of Riemann problems is not considered in the
scheme, given that each interface is treated in an isolated way
4. At each cell i, determine the new cell-averaged values at time level k + 1
averaging the solutions Ui+1/2 and Ui−1/2, e.g.,
Z2Dx Z0
1
1 1
Uki þ 1 ¼ Ui1=2 ðx; DtÞdx þ Ui þ 1=2 ðx; DtÞdx: ð9:30Þ
Dx Dx
0 12Dx
Note that x is the local coordinate for the interface in question. The updated
cell-averaged conservative variables at the new time level are thus a spatial average
of the solution of two Riemann problems solved locally at the two interfaces of each
finite volume. Note that the solution of each Riemann problem is evaluated at time
Δt, and that each solution is determined within half of the cell width. For this
reason, interaction of the waves from the Riemann problems is prohibited and CFL
cannot exceed 1/2 in this version of the scheme.
5. Go to step 1, and repeat the computational cycle until reaching the final target
time.
Figure 9.7 presents schematically a snapshot of computational cell i after time
Dt. Within it, three waves are propagated: a right-going shock generated at interface
i − 1/2 and two left-going shocks coming from interface i + 1/2. The value of
CFL\1=2, so that waves are not interacting within the cell. This simple case
illustrates how to evaluate Eq. (9.30); solving the integrals at time level k + 1 from
point A to B (Fig. 9.7), one finds
1
Uki þ 1 Dx ¼ Ui1=2 SRi1=2 Dt þ Uki Dx SRi1=2 Dt
2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Integration of Riemannproblem from i1=2
1
þ Ui SLi þ 1=2 Dt Dx þ Ui þ 1=2 SRi þ 1=2 Dt SLi þ 1=2 Dt Ukiþ 1 SRi þ 1=2 Dt :
k
2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Integration of Riemann problem from i þ 1=2
ð9:31Þ
360 9 Finite Volume Methods
Fig. 9.7 Riemann solutions at cell i for CFL\1=2: with left-going rarefaction and right-going
shock generated at interface i – 1/2, and two left-going shocks generated at interface i + 1/2.
a Free surface profile at local time Dt, b x-t plane
As shown with this simple example, the updating process may be tedious, espe-
cially if a rarefaction wave is present in the solution at time level k + 1. Further, the
restrictive stability condition CFL\1=2 is another disadvantage of this version of
Godunov’s scheme. It is thus not further considered.
Applying Godunov’s scheme based on Eq. (9.15), the numerical fluxes are
and
Fi þ 1=2 ¼ F Ukiþ 1 : ð9:33Þ
Dt h k
i
Uki þ 1 ¼ Uki F Ui þ 1 F Ui1=2 : ð9:34Þ
Dx
As shown, Eq. (9.34) is simpler to apply than Eq. (9.31), and the maximum time
step Dt could be multiplied by two.
9.3 Approximate Riemann Solvers 361
ZxR ZT
½Uðx; T Þ Uðx; 0Þdx þ ½FðxR ; tÞ FðxL ; tÞdt ¼ 0: ð9:35Þ
xL 0
For the simplified HLL wave structure, Eq. (9.35) yields after elementary
integration
ZxR
Uðx; T Þdx ¼ xR UR xL UL þ T ðFL FR Þdt: ð9:36Þ
xL
ZxR ZTSR
Uðx; T Þdx ¼ ðTSL xL ÞUL þ ðxR TSR ÞUR þ Uðx; T Þdx: ð9:38Þ
xL TSL
362 9 Finite Volume Methods
ZTSR
Uðx; T Þdx ¼ T ðSR UR SL UL þ FL FR Þ; ð9:39Þ
TSL
ZTSR
1 SR UR SL UL þ FL FR
UHLL ¼ Uðx; T Þdx ¼ : ð9:40Þ
T ð SR SL Þ SR SL
TSL
Both the rarefaction waves and shocks are discontinuous waves in the HLL
approximation (see example in Fig. 9.8b). Thus, the Rankine–Hugoniot jump
conditions apply across both, resulting in
FHLL FL
¼ SL ; ð9:41Þ
UHLL UL
and
FHLL FR
¼ SR : ð9:42Þ
UHLL UR
Using the value of UHLL in either of the two jump conditions, the HLL numerical
flux is
SR FL SL FR þ SR SL ðUR UL Þ
FHLL ¼ : ð9:43Þ
SR SL
The possible cases for the numerical flux are thus (Toro 2001)
8
>
> FL if SL 0
>
<
SR FL SL FR þ SR SL ðUR UL Þ
Fi þ 1=2 ¼ if SL 0 SR : ð9:44Þ
>
> SR SL
>
:
FR if SR 0
364 9 Finite Volume Methods
Fig. 9.10 Example of possible wave cases in HLL approximate Riemann solver [corresponding
to Eq. (9.44)]. a Two right-going waves, b left- and right-going waves, c two left-going waves
These are sketched in Fig. 9.10. To apply Eq. (9.44), a reliable estimate of both SR
and SL is needed. For wet-bed conditions, these are given as used in the exact
Riemann solver by (Toro 2001)
SL ¼ U L c L k L ; SR ¼ U R þ c R kR : ð9:45Þ
Here, c = (gh)1/2 is the celerity, and the corrector factor to distinguish between the
celerity of shock propagation, and that of a rarefaction wave given by the corre-
sponding eigenvalue, is (K = L, R)
(h
i1=2
h ðh þ hK Þ
kK ¼
1
2 h2K
h [ hK ðshock wave on K Þ; ð9:46Þ
1 h hK ðrarefaction wave on K Þ:
9.3 Approximate Riemann Solvers 365
Obviously, an estimate of the flow depth in the star region of the exact Riemann
problem is needed. The two-rarefaction result is used here following Toro (2001)
[Eq. (8.22)]
2
1 1 1
h ¼ ð cL þ cR Þ þ ð U L U R Þ : ð9:47Þ
g 2 4
SL ¼ UL cL ; SR ¼ UL þ 2cL ; ð9:48Þ
SL ¼ UR 2cR ; S R ¼ U R þ cR : ð9:49Þ
The Lax diffusive method (Lax 1954; Lax and Wendroff 1960) is a finite-difference
scheme with artificial viscosity (see Chap. 7). It is a variant of the unstable scheme,
given by the updating formula (Cunge 1975; Cunge et al. 1980)
Uki1 þ Ukiþ 1 Dt k
Uki þ 1 ¼ Fi þ 1 Fki1 : ð9:50Þ
2 2Dx
@U @F 1 @ 2 U
þ ¼ D ; ð9:51Þ
@t @x 2 @x2
ðDxÞ2
D¼ : ð9:52Þ
Dt
366 9 Finite Volume Methods
An issue of the Lax scheme is its inconsistency, given that a source term with
artificial viscosity is added to obtain numerical stability. The integral solution of
Eq. (9.51) will be worked out as follows; rewrite it using basic calculus as
@U @ 1 @U
þ F D ¼ 0; ð9:53Þ
@t @x 2 @x
@U @f
þ ¼ 0: ð9:54Þ
@t @x
Its integral form over a control volume in the x-t plane is thus
Dt
Uki þ 1 ¼ Uki f i þ 1=2 f i1=2 : ð9:55Þ
Dx
tZþ Dt
1
f i þ 1=2 ¼ f xi þ 1=2 ; t dt
Dt
t
tZþ Dt
1 1 @U
¼ F D dt
Dt 2 @x i þ 1=2
t ð9:56Þ
tZþ Dt tZþ Dt
1 1 1 @U
¼ F xi þ 1=2 ; t dt D dt
Dt Dt 2 @x
t t
1 @U
¼ Fi þ 1=2 D :
2 @x i þ 1=2
Approximations are now introduced into the general Eq. (9.56). The numerical flux
Fi+1/2 is simply approximated by averaging the physical flux function evaluated at
the initial states of the Riemann problem at interface i + 1/2, whereas the derivative
of U is approximated by a finite difference, resulting in
k
Fki þ Fkiþ 1 1 U Uki Fk þ Fkiþ 1 1 Dx k
f i þ 1=2 ¼ D iþ1 ¼ i Ui þ 1 Uki :
2 2 Dx 2 2 Dt
ð9:57Þ
9.3 Approximate Riemann Solvers 367
Developing the corresponding expression for fi−1/2, and inserting the results into
Eq. (9.55), Eq. (9.50) is regained. In the finite volume interpretation of the Lax
scheme, the solution of the local Riemann problems is avoided by adding artificial
viscosity, thereby forcing numerical stability. The numerical flux of the Lax method
is usually defined as (Toro 2001)
FL þ FR 1 Dx
Fi þ 1=2 ¼ ðUR UL Þ; ð9:58Þ
2 2 Dt
to be used into
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 : ð9:59Þ
Dx
The price of avoiding the solution of the local Riemann problems is to add artificial
viscosity. If the artificial viscosity term is dropped, the numerical flux is
Fki þ Fkiþ 1
Fi þ 1=2 ¼ ; ð9:60Þ
2
resulting in the unstable formula (Hoffman 2001; Hirsch 1988, 1990; LeVeque 2002)
Dt
Uki þ 1 ¼ Uki ðFi þ 1 Fi1 Þ: ð9:61Þ
2Dx
Thus, not solving the Riemann problem and taking the numerical flux as an average
of estimates based on initial data is not a good choice.
An alternative interpretation of Eq. (9.58) is given by Toro (2001) as follows.
Consider the HLL numerical flux given by Eq. (9.44) with the signal speeds
approximated as
SR ¼ SL ¼ Smax ; ð9:62Þ
Dx
Smax ¼ ; ð9:64Þ
Dt
368 9 Finite Volume Methods
which inserted into Eq. (9.63) yields Eq. (9.58). Therefore, the Lax–Friedrichs
scheme can be alternatively understood as a Godunov-type method with the HLL
Riemann solver using the simplest possible choice for the signal speeds.
Roe (1981) developed an approximate Riemann solver for the Euler equations later
applied by Glaister (1987) to the SWE. The homogeneous version of Eq. (9.1) can
be written as
@U @U
þ AðUÞ ¼ 0; ð9:65Þ
@t @x
@F
A¼ : ð9:66Þ
@U
@U ~ @U
þA ¼ 0: ð9:67Þ
@t @x
This matrix is defined locally in terms of the initial data of the Riemann problem at
a generic interface, e.g., A~ ¼A ~ ðUL ; UR Þ: Using the solution to the Riemann
problem for the linear system and resorting to the integral form of the conservation
laws result in the Roe numerical flux (Toro 2001, 2009; Khan and Lai 2014)
FL þ FR 1 X 2
i þ 1=2 ¼
FRoe aj
~
~ kj
K
~ j: ð9:68Þ
2 2 j¼1
~k1 ¼ U
~ ~c; ~k2 ¼ U
~ þ ~c; ð9:70Þ
9.3 Approximate Riemann Solvers 369
Dh ¼ hR hL ; DU ¼ UR UL ; ð9:72Þ
The Roe solver as stated above is not entropy satisfying. Thus, an entropy fix is
needed to avoid unphysical jumps in transcritical rarefactions.
Computationally we must decide whether cells are wet or dry, even though con-
ceptually they could be partially filled with water. This is a practical case when
solving the SWE, given that the position of a shoreline usually lies between cell
centers. Thus, cells can be fully wet, partially wet (partially dry) or dry, but
computationally we must define only wet and dry cells (Brocchini and Dodd 2008).
If the numerical flux is computed using the exact Riemann solver, the dry-bed
cases are automatically considered. The HLL Riemann solver previously exposed
also applies to dry-bed conditions if the correct signal speeds SL and SR for the
dry-bed problem are accounted for, as given by Eqs. (9.48) and (9.49). This permits
reasonably well to compute the numerical flux for dry-bed problems (Toro 2001).
To identify a dry bed in the computational domain, consider a small threshold depth
e (Sanders 2001; Khan and Lai 2014). For a generic interface i + 1/2, the following
cases are possible:
(i) If hL [ e and hR e, the numerical flux Fi+1/2 is computed assuming a dry
bed on the right
(ii) If hR [ e and hL e, the numerical flux Fi+1/2 is computed assuming a dry
bed on the left
(iii) If hR e and hL e, the numerical flux Fi+1/2 is set to zero.
370 9 Finite Volume Methods
After each time loop, a check of computed cell-averaged water depths hki þ 1 is
conducted to identify the formation of new dry-bed conditions within the compu-
tational domain. Two methods are used in practice:
(a) Dry nodes are defined with zero water depth: The following conditions are
implemented in the code
In this method, the depth is clearly zero at the dry nodes. Thus, physical functions
divided by h must be also checked in the code to reset the values to zero under
dry-bed conditions, e.g., the function q2/(gh) or the friction slope Sf = n2q2/(h10/3),
with n as Manning’s roughness coefficient.
(b) Dry nodes are not defined with zero water depth: The following conditions are
implemented in the code
This method can generate unphysical fluxes over uneven beds, so that method (a) is
preferred (Khan and Lai 2014). A typical value used is e = 10–16 m.
Care is needed during assignation of U values at the interface of a wet and a dry
cell, e.g., during the tracking of a shoreline. Suppose an interface with a wet cell on
the left and a dry cell on the right, that is, the shoreline is inside the (computa-
tionally) dry cell, which is in reality a partially-filled cell. Now assume water is
static on a slope. At this interface hL 6¼ 0 from the wet cell, but hR = 0 from the dry
cell, thereby generating unphysical numerical flux. Thus, in this event hR, is set
equal to hL and qR ¼ qL to avoid unphysical flux (Bradford and Sanders 2002).
The problem posed now is the solution of the SWE with a source term S = S(U),
e.g.,
@U @F
þ ¼ SðUÞ: ð9:79Þ
@t @x
9.5 Source Terms 371
U jU j
Sf ¼ Cf ; ð9:80Þ
gh
so that the source term of the SWE is
! !
0 0
S ¼ SðUÞ ¼ @zb ¼ @zb
U 2
U 2
; ð9:81Þ
gh ghSf gU1 Cf
@x @x U 1
U 1
where U = (U1, U2) = (h, q). This source term is reaction-like, given that terms
with U derivatives are missing. Here, a splitting scheme following Toro (2001) is
presented. An alternative to splitting is up-winding the source terms (Bermudez and
Vazquez-Cendón 1994). Our problem is thus to find the vector U at time level k + 1
including the effect of the source terms, that is,
9
@U @F =
PDEs: þ ¼ SðUÞ
@t @x ) Uki þ 1 xi ; tk þ 1 : ð9:82Þ
Initial cond:: Uðx; tÞ ¼ U k ;
i
In the splitting approach, the problem given by Eq. (9.82) is solved in two con-
secutive steps as follows:
Step 1 Solve the homogeneous part of Eq. (9.82) using the Godunov-type scheme,
9
@U @F = k þ 1
PDEs: þ ¼0
@t @x ) Uadv xi ; t : ð9:83Þ
Initial cond:: Uðx; tÞ ¼ Uki ;
i
In this first step, the effects of advection are accounted for. This step may be
regarded as a predictor part to compute an approximate solution at time level k + 1
overlooking source terms effects. From the viewpoint of open channel hydraulics,
this advection step forces the so-called “pseudo-uniform flow condition” (bed-slope
effects in equilibrium with friction). The solution of the advection step is given by
Dt
Uadv ¼ Uki Fi þ 1=2 Fi1=2 ; ð9:84Þ
i
Dx
Step 2 Update the solution including the effect of the source terms,
9
dU =
ODEs: ¼ SðUÞ
dt ) Uki þ 1 xi ; tk þ 1 : ð9:85Þ
Initial cond:: Uðx; tÞ ¼ Uadv ;
i
A variety of ODE solvers apply in this step. It may be regarded as a correction step
to deviate the advection solution from pseudo-uniformity. An integral form of the
ODEs is
tZþ Dt
kþ1
Ui ¼ Ui þ
adv
Si ðUÞdt: ð9:86Þ
t
This equation highlights that the vector Si(U) may be evaluated at different states
during the time integration depending on how the integral is discretized. This fact is
important and may affect the quality of the computed solution. Obviously, this
source term Si(U) shall be regarded as a cell-averaged value.
There is a great variety of ODE solvers to be used in Eq. (9.86); see, for example,
Hoffman (2001). Here, the most widely employed are summarized.
First-order forward Euler scheme
The integral in Eq. (9.86) is evaluated taking the value of S(U) at a specified time
level, resulting in an explicit scheme. Using the value of S(Uk) yields the not best
possible choice (Toro 2001)
k
Uki þ 1 ¼ Uadv
i þ S Ui Dt: ð9:87Þ
A better approximation is
adv
Uki þ 1 ¼ Uadv
i þ S Ui Dt; ð9:88Þ
ð1Þ adv
Ui ¼ Uadv
i þ S Ui Dt; ð9:90Þ
1 1 ð1Þ 1 h ð1Þ i
Uki þ 1 ¼ Uadv þ U þ S Ui Dt: ð9:91Þ
2 i 2 i 2
ð1Þ adv
Ui ¼ Uadv
i þ S Ui Dt; ð9:92Þ
The discretization applied to the bed-slope source term has an important impact on
the quality of the steady-state solutions generated by the unsteady flow solver, an
issue demonstrated as follows. Consider static water over a variable bed profile
zb = zb(x) (Fig. 9.11), for which the conservation laws reduce to
@F
¼ S; ð9:95Þ
@x
or
@ gh2 @zb
¼ gh : ð9:96Þ
@x 2 @x
or,
h2i þ 1=2 h2i1=2 @zb
g ¼ Dx gh : ð9:98Þ
2 @x i
To preserve this identity, the average source term contribution within a cell must be
compatible with the discretization scheme of the numerical fluxes. Otherwise,
artificial waves originating from the numerical scheme may appear, breaking down
static conditions. Topography is defined at cell interfaces, and within a cell, the bed
profile is assumed to vary linearly. The bed source term may be discretized as
@zb hi þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2
gh ¼g : ð9:99Þ
@x i 2 Dx
h2i þ 1=2 h2i1=2 hi þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2
g ¼ g : ð9:102Þ
2 2
9.5 Source Terms 375
Inserting Eqs. (9.99) and (9.102) into Eq. (9.98), the identity is preserved, so the
scheme is said to be well-balanced, or to satisfy the C-property (Bermudez and
Vazquez-Cendón 1994). As noted in this illustrative example, the hydrostatic forces
(gh2/2) contained inside the flux gradient @F=@x are responsible for the potential
imbalance with the bed-slope source term if Eq. (9.99) is not used.
For partially filled cells, e.g. with static water on a slope, the bed jump zbi þ 1=2 zbi1=2
in Eq. (9.99) shall be modified (Brufau et al. 2002). If hi þ 1=2 ¼ 0 then zbi þ 1=2
zbi1=2 is replaced by hi1=2 , whereas for hi1=2 ¼ 0 it is substituted by hi þ 1=2 .
An alternative to circumvent this problem is to reformulate the SWE as follows
@U @F
þ ¼ S; ð9:103Þ
@t @x
where U is the vector of the conserved variables, F the flux vector, and S the source
term vector, given in this case by (Ying et al. 2004)
!
h hU 0
U¼ ; F¼ ; S¼ @zs : ð9:104Þ
hU hU 2 gh ghSf
@x
As noted, the hydrostatic forces are no more contained in F; in the source term,
the bed-slope source term is substituted by a free surface elevation
(zs = zb + h) gradient term. For static water, the integral form of Eq. (9.103)
reduces to
@zs
Dx gh ¼ 0: ð9:105Þ
@x i
Consider the bed-slope source term overlooked temporarily. In this case, and with
f as the Darcy–Weisbach friction factor (Cf = f/8) (White 2009), Eq. (9.85) is
reduced to the scalar ODE, assuming a wide rectangular channel and positive U
dq f
¼ ghSf ¼ U 2 : ð9:106Þ
dt 8
However, near wet–dry fronts the explicit discretization may generate instabilities,
so that an alternative method is explained here. Using backward Euler discretization
[Eq. (9.89)], Eq. (9.106) yields
f k þ 1 2
qki þ 1 ¼ qadv
i U Dt; ð9:108Þ
8 i
or
qadv f k þ 1 2
Uik þ 1 ¼ i
U Dt: ð9:109Þ
hki þ 1 8hki þ 1 i
Equation (9.109) is a quadratic function solved at each time step to obtain the
cell-averaged velocity Uk+1
i accounting for frictional effects. Note that the friction
force might stop the flow in the extreme; e.g., a flow reversal cannot be numerically
permitted. Thus, the following condition must be verified by numerical
computations
qki þ 1 qadv
i 0: ð9:110Þ
Other flow resistance formulae like Manning’s equation are frequently used in river
flow computations (Henderson 1966; Cunge et al. 1980; Khan and Lai 2014).
A semi-implicit approach is also used in some models.
In this section, the first-order upwind finite volume method of Ying et al. (2004) is
described. In this model, the SWE are used in the form given by Eqs. (9.103) and
(9.104), so the scheme is well-balanced. The updating formula is given by the
integral equation
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 þ DtSi : ð9:111Þ
Dx
The numerical flux is evaluated with the one-sided upwind method as (Ying et al.
2004)
2 3
qkiþ j
6 k 27
Fi þ 1=2 ¼6
4 qi þ j 5 :
7 ð9:112Þ
hkiþ j
9.6 One-Sided First-Order Upwind Scheme 377
Here j = 0 if qi and qi+1 > 0, j = 1 if qi and qi+1 < 0, and j = 1/2 for any other case;
subscript i + 1/2 means in that case average values between i and i + 1
cell-averaged values. The water surface gradient cannot be evaluated with a central
difference, given that non-physical jumps are then formed within rarefaction waves
(Ying et al. 2004). Thus, an average between up- and downwind gradients is
formed as (Ying et al. 2004)
!
@zs kþ1
zks;iþþ11j zks;ij
þ1
zks;iþþ1j zks;i1
þ1
þj
gh ¼ ghi w1 þ w2 : ð9:113Þ
@x Dx Dx
Note that the water depths used in Eq. (9.113) are those at time k + 1, after solving
the advection step for the continuity equation. The weighting factors are
The computational sequence to apply the scheme encompasses the following steps:
(1) Start with known cell-averaged values of U at a time level k.
(2) Determine a stable time step Dt using the CFL condition, e.g., typically with
CFL ¼ 0:9, using Eq. (9.29).
(3) Compute the numerical fluxes Fi+1/2 using Eq. (9.112).
(4) Compute the water depths solving the advection step for the continuity
equation
Dt
hki þ 1 ¼ hki qi þ 1=2 qi1=2 : ð9:115Þ
Dx
(5) Evaluate the water surface gradient source term using Eq. (9.113).
(6) Compute the discharge at the new time level including the effects of source
terms as
qki þ 1 ¼ qki
" # !
Dt q2i þ j q2ij kþ1
zks;iþþ11j zks;ij
þ1
zks;iþþ1j zks;i1
þ1
þj
ghi w1 þ w2 Dt
Dx hi þ j hij Dx Dx
f
Uik
Uik
Dt:
8
ð9:116Þ
(7) Go to step (1), and repeat the cycle until reaching the final time of
computations.
378 9 Finite Volume Methods
van Leer introduced a second-order version of the Godunov (1959) scheme (van
Leer 1979). The scheme is called MUSCL (Monotonic Upstream Centered
Schemes for Conservation Laws). Second-order accuracy is regained by recon-
structing the solution U(x) within the cells. The idea is to replace data representation
using the constant cell-averaged value Ui by a function U(xi−1/2 < x < xi+1/2) that
approximates the exact solution within a cell. The function used for the recon-
struction must preserve the cell-averaged value regained from the conservative
formula. That is, the latest cell-averaged values available obtained by applying the
time stepping are used in the reconstruction of the solution at the actual time level.
In van Leer’s scheme, the reconstruction of the solution within a cell is linear
(Fig. 9.12a, b), but parabolic approximations also apply (Colella and Woodward
1984). The linear reconstruction of the solution U(x) within a cell gives
DUi
Ui ð xÞ ¼ Uki þ ðx xi Þ : ð9:117Þ
Dx
Note that the slope vector is constant for a given cell, but each conserved variable
will have a different scalar slope contained therein. To compute the slope vector for
the reconstructed solution, DUi =Dx; up- and downstream jumps are first defined as
Taking an average slope, with the averaging coefficient −1 < x < 1, produces
(Toro 2001)
1
DUi ¼ ð1 xÞDUi þ 1=2 þ ð1 þ xÞDUi1=2 : ð9:119Þ
2
A typical value is x = 0. The boundary extrapolated values at each cell face are
thus (Fig. 9.12b)
1 1
Uki1=2 ¼ Uki DUi ; Ukiþ 1=2 ¼ Uki þ DUi : ð9:120Þ
2 2
1 1
ULiþ 1=2 ¼ Uki þ DUi ; URiþ 1=2 ¼ Ukiþ 1 DUi þ 1 : ð9:121Þ
2 2
9.7 MUSCL-Hancock Second-Order TVD Scheme 379
Bram van Leer was born on November 26, 1942, at Surabaya, Netherlands East
Indies. He obtained the Ph.D. degree from Leiden State University in 1970. He
made seminal contributions to computational fluid dynamics (CFD) in his 5-part
article series “Towards the ultimate conservative difference scheme” published
from 1972 to 1979, where he extended Godunov’s finite volume scheme to
second order (MUSCL), developed non-oscillatory interpolation using limiters, an
approximate Riemann solver, and discontinuous Galerkin schemes for unsteady
advection. Since joining the University of Michigan’s Aerospace Engineering
Department in 1986, he has worked on convergence acceleration by local pre-
conditioning and multigrid relaxation for Euler and Navier–Stokes problems,
unsteady adaptive grids, space-environment modeling, atmospheric flow model-
ing, extended hydrodynamics for rarefied flows, and discontinuous Galerkin
methods. He retired in 2012. His research interests were in CFD, fluid dynamics,
and numerical analysis.
He obtained in 1996 the College of Engineering Research Award
(University of Michigan), was elected in 1995 AIAA Fellow, was awarded in
1990 and 1992 the NASA Langley Group Achievement Award, in 1992 the
Department of Aerospace Engineering Research Award (University of
Michigan), the Honorary Doctorate from the Vrije Universiteit Brussels in
1990, and the C. J. Kok Prize of Leiden State University in 1978. He is a
Fellow of the American Institute of Aeronautics and Astronautics (AIAA)
and Member of the Society for Industrial and Applied Mathematics (SIAM).
using the so-called total variation diminishing (TVD) schemes. They suppress
Gibbs’ phenomenon by degradation of the scheme to first-order accuracy in zones
with sharp variations of U, while keeping second-order accuracy in smooth portions
of the solution. Spurious oscillations are suppressed by limiting the slope in the
reconstruction stage. Thus, the reconstruction of the solution reads now with
DUi =Dx as the limited slope vector
DUi
Ui ð xÞ ¼ Uki þ ðx xi Þ ;
Dx ð9:122Þ
1 1
ULiþ 1=2 ¼ Uki þ DUi ; URiþ 1=2 ¼ Ukiþ 1 DUi þ 1 :
2 2
The idea is to limit the slopes obtained in the reconstruction stage by applying a
limiter function. A widely used limiting function is the so-called minmod limiter,
which for two numbers a and b is
The minmod function gives the smallest modulus if all arguments have the
same sign and is zero otherwise. For the MUSCL scheme, the minmod limiter
is thus
DUi ¼ sign DUi þ 1=2 max 0; min
DUi þ 1=2
; sign DUi þ 1=2 DUi1=2 :
ð9:125Þ
The three possible cases are depicted in Fig. 9.13. If the jumps have the same sign
(Fig. 9.13a, b), the solution is monotone. Therefore, the limited jump is nonzero, and
equal to the smaller one in absolute value. If the jumps change their signs, the up- and
downslopes have different signs (Fig. 9.13c), thereby implying a solution U(x) of
certain curvature (oscillation). The propagation of these non-physical undulations is
constrained turning to first-order accuracy the scheme by setting the limited jump to
zero. Thus, the linear reconstruction is not done in these cases. Therefore, the TVD
scheme is adaptive; it decides the regions of the computational domain where the
scheme has to be second-order accurate and those where the linear reconstructions must
382 9 Finite Volume Methods
Fig. 9.13 Effect of minmod slope limiter. a Limited jump is the upstream jump, b Limited jump
is the downstream jump, and c limited jump is zero [the figure represents a component of the
vector U = (U1, U2)T]
DUi
(
sign DUi þ 1=2 min min
DUi þ 1=2
;
DUi1=2
; b min
DUi þ 1=2
;
DUi1=2
if DUi þ 1=2 DUi1=2 [ 0
¼
0 else:
ð9:126Þ
The MUSCL scheme (van Leer 1979) described is second-order accurate in space,
but only first-order accurate in time. It was simplified and improved by Hancock
(1980) while developing the PISCES industrial simulation code (Van Leer 2006).
The method would have remained unnoticed in the manual of this code, but it was
rescued by van Albada et al. (1982) and applied for the simulation of cosmic gas
dynamics. The idea of Hancock’s predictor–corrector scheme is remarkably simple:
A first-order upwind method transforms into second-order time accuracy by
evolving the cell boundary values used to compute the numerical flux, and the
source terms, by half the time step (Sweby 1999). Thus, to regain second-order
accuracy in both time and space, the vector U is expanded using a Taylor series
k þ 1=2
within a cell as follows. The cell interface value evolved a time Dt/2, Ui þ 1=2 ¼
1
U 2 Dx; 12 Dt ; can be expressed as a function of the cell-averaged value at time
k using a Taylor series for two variables as (Sweby 1999)
k k
k þ 1=2 1 @U 1 @U
Ui þ 1=2 ¼ Uki þ Dx þ Dt þ O Dx2 ; Dt2 : ð9:127Þ
2 @x i 2 @t i
Note that the value of U at the cell center at time k is the cell-averaged value, as
assured by the MUSCL linear reconstruction at time level k. The time derivative of
U is from the differential form of the SWE
@U @F
¼ þ S: ð9:128Þ
@t @x
k F Uk
i þ 1=2 F Ui1=2
k
@F
; ð9:131Þ
@x i Dx
384 9 Finite Volume Methods
the relations needed in Eq. (9.129), namely Eqs. (9.130)–(9.132), are available.
Inserting these, one gets
1 U k
i þ 1=2 U k
1 F Uk
i þ 1=2 F Uk
i1=2 1
k þ 1=2 i1=2
Ui þ 1=2 ¼ Uki þ Dx Dt þ DtS Uki
2 Dx
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} 2 Dx 2
MUSCL reconstruction
Dt h k
i Dt
¼ Ukiþ 1=2 F Ui þ 1=2 F Uki1=2 þ S Uki :
2Dx 2
ð9:133Þ
This is Hancock’s step to regain second-order time accuracy: The MUSCL reconstructed
interface values Ukiþ 1=2 at time k are evolved half the time step to obtain an improved
k þ 1=2
data representation given by Ui þ 1=2 . Looking to an interface rather than to a single cell,
Hancock’s step amounts to evolve the interface values ULi þ 1=2 and URi þ 1=2 to regain second-
L R
order accuracy in time to Ui þ 1=2 and Ui þ 1=2 (Fig. 9.14a). These are given by Hancock’s
scheme from Eq. (9.133) as
L Dt h L
i Dt
Ui þ 1=2 ¼ ULi þ 1=2 F Ui þ 1=2 F URi1=2 þ Si ;
2Dx 2 ð9:134Þ
R Dt h L
i Dt
Ui þ 1=2 ¼ Ui þ 1=2
R
F Ui þ 3=2 F URi þ 1=2 þ Si þ 1 ;
2Dx 2
forming a new local Riemann problem. With these evolved boundary extrapolated
L R
variables Ui þ 1=2 and Ui þ 1=2 defining states L and R, the numerical flux is computed
using the HLL approximate Riemann solver (Fig. 9.14b). Once this flux is deter-
mined, the computation of the transient flow is conducted as described previously.
Note that this predictor step considers the source terms. The bed-slope effect is
typically accounted for (Zhou et al. 2001; Aureli et al. 2008). It is evaluated based
on interface data at cell i as [see Eq. (9.99)]
hL R
@zb i þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2
gh ¼g : ð9:135Þ
@x i 2 Dx
Friction is treated in a final splitting step using an implicit treatment (Aureli et al.
2008) to increase stability in the vicinity of wet–dry interfaces.
Dt h
L R
L R
i
Uadv ¼ Uki Fi þ 1=2 Ui þ 1=2 ; Ui þ 1=2 Fi1=2 Ui1=2 ; Ui1=2 : ð9:136Þ
i
Dx
386 9 Finite Volume Methods
(8) Include the effect of the bed-slope source terms using the values obtained at
time k + 1/2 with Hancock’s method, e.g.,
T
~ i ¼ Uadv 0; Dt g hL
U þ h
R
z bi þ 1=2 z : ð9:137Þ
i i þ 1=2 i1=2 bi1=2
2Dx
Combining Eqs. (9.136) and (9.137) yields (Mingham and Causon 1998; Aureli
et al. 2008)
h
i
~ i ¼ Uk Dt Fi þ 1=2 UL
U ; U
R
F U
L
; U
R
i þ 1=2 i þ 1=2 i1=2
i
Dx i1=2 i1=2
T ð9:138Þ
Dt L R
0; g hi þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2 :
2Dx
This is only a function of the predicted values at k + 1/2 obtained with Hancock’s
method.
(9) Include the effect of friction using an implicit Euler scheme
T
~ i Dt 0; f U k þ 1
U k þ 1
:
Uki þ 1 ¼ U ð9:139Þ
8 i i
The MUSCL reconstruction process of the vector U = (h, q)T (step 3 in the
sequence above) is called depth gradient method (DGM) (Zhou et al. 2001); it may
lead to unphysical discharges over uneven topography; given that for static water,
the reconstruction of variable h will produce jumps in water depths at cell interfaces
and thus a water movement even if the water is initially at rest. Figure 9.15 depicts
the application of DGM to an uneven profile, from where is seen that
A linear reconstruction yields a local Riemann problem at each cell involving left
and right states
9.7 MUSCL-Hancock Second-Order TVD Scheme 387
Fig. 9.15 Generation of jump in water depths at cell interfaces by the depth gradient method over
uneven topography
! !
hLiþ 1=2 hRiþ 1=2
ULiþ 1=2 ¼ ; URiþ 1=2 ¼ : ð9:141Þ
qLiþ 1=2 ¼ 0 qRiþ 1=2 ¼ 0
Assume to compute now the first time step of the transient flow, starting with zero
discharge in the system and a horizontal free surface. Reconstruction of the depth
profile yields the momentum equation
2
2
< <
Dt Fi þ 1=2 Fi1=2 hLiþ 1=2 þ hRi1=2 zbi þ 1=2 zbi1=2
qki þ 1 ¼ g gDt 6¼ 0:
Dx 2 2 Dx
ð9:142Þ
In this equation, Fi<þ 1=2 is the momentum function obtained at the interface i + 1/2
solving the Riemann problem formed there. Obviously, the discharge is in general
not zero, indicating the generation of an unphysical discharge. If
hLiþ 1=2 ¼ hRiþ 1=2 ¼ hi þ 1=2 , Eq. (9.142) reduces to
h 2
Dt i þ 1=2 h 2
i1=2 hi þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2
qki þ 1 ¼ g gDt ¼ 0;
Dx 2 2 Dx
ð9:143Þ
1 1
Qki1=2 ¼ Qki DQi ; Qkiþ 1=2 ¼ Qki þ DQi ; ð9:144Þ
2 2
where
zs zb þ h
Q¼ ¼ : ð9:145Þ
q q
Once the interface values are determined with the limited slopes, the water depths at
the cell faces are
hki1=2 ¼ ðzs Þki1=2 ðzb Þi1=2 ; hkiþ 1=2 ¼ ðzs Þkiþ 1=2 ðzb Þi þ 1=2 : ð9:146Þ
With this technique, the identity given by Eq. (9.143) is preserved, and thus,
unphysical discharge is not generated, given that the flow depths at the cell faces are
unique for static conditions. This technique is referred to as the surface gradient
method (SGM). This reconstruction, however, can lead to small, even negative
water depths near dry–wet fronts (Aureli et al. 2008). DGM, in contrast, is more
robust for bore front tracking. Aureli et al. (2008) propose a hybrid reconstruction
combining the good capabilities of the SGM and DGM, called weighted
surface-depth gradient method.
The computational domain is typically divided into N cells (Fig. 9.16a) of width
Δx, e.g., with the initial data
Uki ; i ¼ 1; 2; . . .. . .. . .. . .; N 1; N: ð9:147Þ
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 þ DtSi ; i ¼ 1; 2; . . .. . .. . .. . .; N 1; N:
Dx
ð9:148Þ
Fig. 9.16 Boundary conditions. a Definition of ghost cells in the computational domain, ghost
cells in (b) weir flow, c gate flow
Dt
Uk1 þ 1 ¼ Uk1 F3=2 F1=2 þ DtS1 : ð9:149Þ
Dx
In this formula, the numerical flux entering into the left face of cell i = 1, namely
F1/2, is needed. However, there is no initial data defined to form a Riemann problem
at this interface and thus compute the numerical flux.
390 9 Finite Volume Methods
Dt
UkNþ 1 ¼ UkN FN þ 1=2 FN1=2 þ DtSN : ð9:150Þ
Dx
In this formula, the numerical flux leaving the right face of cell i = N, namely FN+1/2,
is needed. However, there is no initial data defined to form a Riemann problem at
this interface and thus compute the numerical flux.
From the above arguments, it is necessary to incorporate two additional cells at
the left and right ends of the computational domain, called ghost cells (LeVeque
2002; Toro 2009) (Fig. 9.16a). Physical boundary conditions are transmitted to the
computational model through these cells, which in turn are used to form two
additional Riemann problems needed to compute the missing numerical fluxes and
thus evolve in time cells i = 1 and N. Here, the simplest way of implementing
boundary conditions is exposed using ghost cells with a zero-order extrapolation
from the interior solution. Care should be taken to ensure that spurious waves are
not generated when implementing boundary conditions (LeVeque 2002).
Subcritical inflow section
At a subcritical inflow section, one boundary condition must be prescribed, as
known from the method of characteristics, explained in Chap. 5 (Henderson 1966;
Jain 2001). This is typically the inlet discharge for weir flow (Fig. 9.16b). The other
variable needed at the ghost cell is obtained approximately by extrapolation from
the interior solution (LeVeque 2002). The vector U at the ghost cell i = 0 is thus
given by
qko þ 1 ¼ qinlet ;
ð9:151Þ
hko þ 1 ¼ hk1 þ 1 :
For a more rigorous computation of hko þ 1 the characteristic C - shall be used (see
Chaps. 5 and 7; Eq. 7.37).
Subcritical outflow section
At a supercritical outflow section, boundary conditions are not needed (see Chap. 5).
The other variable required at the ghost cell is obtained approximately by extrapolation
from the interior solution (LeVeque 2002). This is typical for supercritical flow at the
tailwater portion of weir flow (Fig. 9.16b). The vector U at the ghost cell i = N + 1 is
thus given by
qkNþþ11 ¼ qkNþ 1 ;
ð9:152Þ
hkNþþ11 ¼ houtlet :
For a more rigorous computation of qkNþþ11 the characteristic C + shall be used (see
Chaps. 5 and 7; Eq. 7.39).
9.8 Boundary and Initial Conditions 391
qko þ 1 ¼ qinlet ;
ð9:153Þ
hko þ 1 ¼ hinlet :
qkNþþ11 ¼ qkNþ 1 ;
ð9:154Þ
hkNþþ11 ¼ hkNþ 1 :
After application of the conservative equation in a given time loop, the following
data are generated:
Uki þ 1 ; i ¼ 1; 2; . . .. . .. . .; N 1; N: ð9:155Þ
Once this step finished, U at the ghost cells is determined depending on the type
of flow section as exposed above and summarized here as:
k þ 1
h
Uko þ 1 ¼
q o
one boundary condition; one value extrapolated from i ¼ 1 if F1 \1
¼
two boundary conditions if F1 [ 1;
k þ 1
h
UkNþþ11 ¼
q N þ1
one boundary condition; one value extrapolated from i ¼ N if FN \1
¼
two values extrapolated from i ¼ N if FN [ 1:
ð9:156Þ
392 9 Finite Volume Methods
The initial conditions are the values of vector U at the computational cells at the
initial instant at which transient computations start, e.g.,
0
h
U0i ¼ ; i ¼ 1; 2; . . .. . .. . .. . .; N 1; N: ð9:157Þ
q i
If the water is initially static over a variable bottom profile of constant elevation zs,
the vector is
0
h zs zbi
U0i ¼ ¼ ; i ¼ 1; 2; . . .. . .. . .; N 1; N: ð9:158Þ
q i 0
This is a wet-bed initial vector. For dry-bed portions of the computational domain,
the vector reads
0
h 0
U0i ¼ ¼ ; i ¼ 1; 2; . . .. . .. . .; N 1; N: ð9:159Þ
q i 0
The water may be also under steady-state conditions, once changes in boundary
conditions are introduced. Then, the initial discharge is a constant qo in the absence
of sources/sinks, e.g., infiltration, side weirs, or precipitation. The water depth
profile is varied, ho = ho(x), and it has to be determined with the steady-state
techniques of Chaps. 3 and 4, or taken from the steady asymptotic state of a
previous unsteady flow computation. The initial vector then reads
0
h ho ¼ ho ð x Þ
U0i ¼ ¼ ; i ¼ 1; 2; . . .. . .. . .. . .; N 1; N: ð9:160Þ
q i qo ¼ const:
The MUSCL-Hancock scheme using the HLL approximate Riemann solver was
coded and implemented in the file DamBreakWave.xls available in Chap. 12. The
code can be run with friction, and the selected value of b in the limiter [Eq. (9.126)]
fixes the use of minmod (b = 1), Superbee (b = 2) or changes the second-order
scheme to the Godunov first-order scheme if b = 0. In this section, ideal fluid flow
computations in a horizontal channel using the minmod slope limiter are conducted.
9.9 Computational Examples 393
If the limiter is deactivated inside the code, spurious oscillations remain unsup-
pressed contaminating numerical solutions.
A subcritical dam break wave of depth ratio hd/hu = 0.4 and upstream water
depth hu = 1 m is considered in Fig. 9.17a, b. Numerical simulations were con-
ducted using CFL ¼ 0:9 and Dx = 0.1 m, with computational results displayed at
time t = 3 s. In the same figure, the exact solution of Stoker (1957), detailed in
Chap. 6, is plotted. The MUSCL-Hancock scheme is shown to produce an excellent
oscillation-free solution of the wave profiles, including the correct shock strength
and position. A most stringent test case is generated in Figs. 9.17c, d using a depth
ratio hd/hu = 0.01, hu = 1 m and identical values of computational parameters.
Critical flow establishes at the dam axis x = 0 in a point of the rarefaction wave. As
depicted in the comparison of the numerical solution with the exact results, the
MUSCL-Hancock scheme produces a highly accurate resolution of transcritical
flow near the dam axis. Therefore, the scheme is entropy preserving and unphysical
jumps are avoided. Finally, a dry-bed dam break wave is considered in Figs. 9.17e, f.
Dry beds were handled in the numerical model using Eqs. (9.76) and (9.77), and
the computational results show good agreement with Ritter’s analytical solution
(Ritter 1892). Note that the dry front in Fig. 9.16e is faster than the shock fronts in
Figs. 9.16a, c. A more detailed study of numerical schemes applied to the dam break
problem is available (Zoppou and Roberts 2003).
U2
Sf ¼ n2 4=3
; ð9:161Þ
Rh
Fig. 9.17 Ideal dam break wave simulations: comparison of numerical simulations using the
MUSCL-Hancock scheme with exact solutions for a, b subcritical wet-bed test, c, d transcritical
wet-bed test, e, f dry-bed test
or
k þ 1 2
qadv 2 Ui
Uik þ 1 ¼ i
gn k þ 1 4=3 Dt: ð9:164Þ
hki þ 1 R hi
friction source term and the minmod slope limiter. During initiation of motion, the
wave profile is affected by non-hydrostatic flow effects (Castro-Orgaz and Chanson
2017), so that precise agreement of experiments with simulations based on the SWE
is not expected. The overall position of the surge front is reasonably predicted, but
the detailed wave shape is affected by turbulence, a feature not accounted for in the
mathematical model. Experiments (Ozmen-Cagatay and Kocaman 2010) and
numerical simulations using the MUSCL-Hancock scheme for a transcritical dam
break wave test (hd/ho < 0.138) are considered in Fig. 9.19, with hd = 0.025 m. For
this test, the position of the surge is reasonably well predicted, but the effects of
turbulence are strong in the supercritical portion of the wave profile. A dry-bed dam
break wave is examined in Fig. 9.20, showing a reasonable agreement of numerical
predictions with experimental data by Ozmen-Cagatay and Kocaman (2010). The
shape of the experimental curves of Fig. 9.20 is similar to those previously mea-
sured by Dressler (1954) or Lauber (1997).
The experiments by Schoklitsch (1917) for a dam break wave in a dry, rect-
angular, horizontal flume are considered as an additional test case. The flume was
0.093 m wide, 0.08 m high, and 20 m long. The dam was located at x = 10 m, and
the removal was considered instantaneous. The tailwater flume portion was initially
dry, and the water depth at the dam 0.074 m. Experimental data recorded by
Schoklitsch (1917) for two times after dam removal, namely t = 3.75 s and
t = 9.4 s, are plotted in Fig. 9.21a, b, respectively. The predictions using the
MUSCL-Hancock scheme were conducted using CFL ¼ 0:9, Dx = 0.1 m, and
n = 0.008 m−1/3s. A comparison of simulations and data shows good agreement.
The effect of flow resistance in this dataset is significant for the wave front, as
observed from the comparison of the numerical simulation with Ritter’s solution
plotted in the same figure.
is detected, given the use of Eq. (9.12). This is purely a numerical issue common to
many numerical schemes (Khan and Lai 2014). A comparison of the water surface
profile with experimental data by Gharangik and Chaudhry (1991) in Fig. 9.23c
shows good agreement, although the finite length of the jump produced by tur-
bulence is not a feature tractable with the numerical model of the SWE.
9.9 Computational Examples 401
An important transcritical open channel flow feature is the passage from sub-
(F\1Þ to supercritical (F [ 1) flow over a round-crested weir (Fig. 9.16b). Here,
this steady transcritical flow is simulated using the MUSCL-Hancock scheme
assuming ideal fluid flow. An initial steady free surface profile over the weir, for
which qo = const. and ho = ho(x) are known, must be prescribed to initiate unsteady
computations [Eq. (9.160)]. Assuming Sf = 0, Eq. (9.1) reduces for steady flow to
(Henderson 1966)
402 9 Finite Volume Methods
dh @zb
Sf @zb
¼ @x 2 ¼ @x
2 : ð9:165Þ
dx 1F 1 gh3
q
In this test, Eq. (9.165) was used to produce an initial free surface profile over the
weir. The profile was numerically computed using the fourth-order Runge–Kutta
method (Chaudhry 2008), as described in Chaps. 3 and 4. Computations started at the
crest section where the flow is critical (F ¼ 1Þ; that is, h = hc = (q2/g)1/3. If F ¼ 1;
then Eq. (9.165) must equal the indeterminate identity dh/dx = 0/0. This singularity
is removed by applying L’Hospital’s rule to Eq. (9.165), resulting in (see Chap. 4)
1=2
dh hc @ 2 z b
¼ : ð9:166Þ
dx c 3 @x2
At the crest section, Eq. (9.166) was implemented in the Runge–Kutta solver, and
the corresponding sub- and supercritical branches of the water surface profile were
computed in the up- and downstream directions, respectively.
For unsteady transcritical flow over a weir, one boundary condition must be
prescribed at the subcritical section on the upstream weir side, whereas no boundary
conditions need to be prescribed at the supercritical outlet section. The inlet
boundary condition is given by an instantaneous rise in the discharge, which is kept
constant during all transient flow. Unknown values of conserved variables at
boundary sections are then computed using ghost cells by extrapolation of values at
adjacent interior cells (LeVeque 2002), as previously explained. The steady water
surface profile over a weir of bed shape zb = 0.2 − 0.01x2 (m) was computed using
the MUSCL-Hancock method with the HLL approximate Riemann solver and the
DGM for reconstruction, for a target discharge of q = 0.2 m2/s. It was settled
instantaneously at the inlet, once transient computations were initiated, and kept
constant during all the routing. The initial free surface profile was computed for the
steady discharge qo = 0.05 m2/s.
This particular weir is widely used to test unsteady numerical models (i.e., Zhou
et al. 2001; Ying et al. 2004; Castro-Orgaz and Chanson 2016). In this test case,
Δx = 0.05 m and CFL ¼ 0:9 were used. The code is available in the file
Weir_DGM_parabolic.xls. In the code Weir_DGM_parabolic_MC.xls, the
upstream water depth is determined by using a backward characteristic, showing
variations at initiation of motion only (t < 0.05 s). Some snapshots of the unsteady
flow motion are presented in Fig. 9.24, from where it is observed how a shock wave
passes over the weir (Fig. 9.24a–c). Once it disappears, e.g., at t = 3 s (Fig. 9.24d),
the unsteady free surface flow profile is smooth and gradually adjusts to the dise-
quilibrium in discharge along the weir profile. The results of Fig. 9.24e involve a
simulation time of t = 350 s to ensure that steady-state conditions are reached. The
steady water surface profiles computed using Eqs. (9.165) and (9.166) are presented
in the same figure (labeled as exact solution), showing the excellent agreement of
the finite volume computations with these results. Note further that the discharge is
well conserved by the unsteady flow model. The same computations conducted
9.9 Computational Examples 403
Fig. 9.24 Unsteady transcritical flow using the MUSCL-Hancock scheme: evolution of bore
passage over a weir profile and comparison of steady-state solution with exact results
using the one-sided upwind finite volume method yield results almost identical
(Castro-Orgaz and Chanson 2016) (the two profiles deviate only in the third digit).
This numerical model is implemented in the file Weir_upwind.xls of Chap. 12.
The code Weir_dry_SGM-DGM_MC.xls permits to simulate this steady flow and
allows for flooding over dry terrain.
Figure 9.25 contains the experimental data of Sivakumaran et al. (1983) for a
Gaussian hump of profile zb = 20exp[− 0.5(x/24)2] (cm) for two test cases. This
weir test is implemented in the file Weir_DGM_Sivakumaran.xls in Chap. 12. The
computed solution using the MUSCL-Hancock scheme is presented for both cases
and compared in Fig. 9.25a, c with the exact steady solution obtainable using
Eqs. (9.165) and (9.166). Steady solutions obtained with the MUSCL-Hancock
404 9 Finite Volume Methods
Fig. 9.25 Transcritical flow over Gaussian weir profile using the MUSCL-Hancock scheme.
Comparison of steady-state solutions with exact results and experimental data (Sivakumaran et al.
1983) for q = a, b 0.111 m2/s, c, d 0.0359 m2/s
scheme are in excellent agreement with exact results (Fig. 9.25a, c). The departure
between simulations and experiments for the test case of Fig. 9.25b (q = 0.111 m2/s)
indicates that the effect of the vertical acceleration as the flow passes from sub- to
supercritical is notable, so that the solution of the SWE does not fully agree with
physical experiments due to non-hydrostatic effects. For the test case of Fig. 9.25d
(q = 0.0359 m2/s), deviations between numerical results and experiments are small.
The SWE produce realistic free surface profile solutions across the critical depth
using shock-capturing numerical methods. The computation of a steady flow profile
using an unsteady flow computation produces a solution that automatically crosses
the critical depth. This unsteady flow computation is performed without any further
special treatment at the critical point, as the unsteady computation does not suffer
from any mathematical indetermination. However, the steady gradually-varied flow
equation has a mathematical indetermination at critical flow conditions resolved
with L’Hopital’s rule. The unsteady computation produces a singular point at the
weir crest automatically as the steady state is asymptotically reached (Castro-Orgaz
and Chanson 2016).
9.9 Computational Examples 405
Consider the experimental tests of Ozmen-Cagatay and Kocaman (2011) for a dam
break wave propagated over a trapezoidal sill. The upstream water depth was
ho = 0.25 m, the flume width was 0.3 m, and the downstream channel was dry.
A trapezoidal bottom sill 0.075 m high, of 0.3 m crest length, with up- and
downstream slopes of 7.5/35, was inserted 1.53 m downstream of the gate used to
simulate the dam section. The upstream flume portion was used to store the water
and simulate a reservoir, over a length of 4.65 m. This test was modeled with the
one-sided first-order finite volume method (Ying et al. 2004). Simulations were
conducted using CFL ¼ 0:9 and Dx = 0.01 m assuming ideal fluid flow. The
downstream boundary conditions were modeled as transmissive boundary condi-
tions (Toro 2001), e.g.,
qkNþþ11 ¼ qkNþ 1 ;
ð9:167Þ
hkNþþ11 ¼ hkNþ 1 :
These are also the conditions used to model supercritical flow at an outlet. The
upstream portion of the computational domain is a solid wall, which must be
modeled using reflective boundary conditions (Toro 2001)
qko þ 1 ¼ qk1 þ 1 ;
ð9:168Þ
hko þ 1 ¼ hk1 þ 1 :
The physical effect of this condition is to represent what actually happens at a solid
wall: There is no mass flow, and there is a hydrostatic thrust acting on the wall.
Instead of using Eq. (9.168) it is possible to fix the numerical flux directly. The
computational model is available in the code of the file “Dambreak_sill.xls”,
Chap. 12. The evolution of this flow is observed in Fig. 9.26, showing the initial
run-up of a supercritical current over the trapezoidal sill, and its subsequent
drowning with wave breaking and bore propagation in the upstream direction. The
experimental data of Ozmen-Cagatay and Kocaman (2011) at different normalized
times T = t(g/ho)1/2 are considered in Fig. 9.27, where computational results at the
same times are also included. Note that the SWE produce an excellent represen-
tation of the supercritical flow passage over the trapezoidal sill and adequately
represent wave breaking and formation of the bore propagating in the upstream
direction at later times. The precise shape of the upstream bore is not well predicted,
given that the SWE are unable to model the finite bore length due to turbulence, but
the overall position of the bore is in good agreement with computations. The
propagation of a supercritical flow over a sill and subsequent wave breaking and
propagation of an upstream bore is a beautiful hydraulic feature (Fig. 9.28) to be
handled with satisfactory results using the SWE.
406 9 Finite Volume Methods
Fig. 9.28 Supercritical flow over a sill resulting in wave breaking and a surge propagating in the
upstream direction (flow is from left to right) [Taken from movie Fluid motion in a gravitational
field, by Rouse (1961), IIHR-Hydroscience & Engineering, The University of Iowa]
1=2 rffiffiffiffiffiffiffiffiffi!
4ho 1
L¼ arcosh ; ð9:169Þ
3H 0:05
where ho = 1 m is the still water depth and H the solitary wave amplitude. A test for
non-breaking solitary wave run-up with H/ho = 0.04 is presented in Fig. 9.29 to
check the predictions of the SWE. The initial condition is given by the solitary
wave solution
" 1=2 #
3H
gðx; 0Þ ¼ Hsech2 ð x þ LÞ ; ð9:170Þ
4h3o
Fig. 9.29 Non-breaking solitary wave run-up on a plane beach for H/ho = 0.04 at various
normalized times T = t(g/ho)1/2
410 9 Finite Volume Methods
1=2
g
Uðx; 0Þ ¼ gðx; 0Þ ; ð9:171Þ
ho
where η = h – ho is the free surface displacement. Solutions were obtained with the
MUSCL-Hancock scheme using an implicit treatment of the friction effects.
Computations were conducted using CFL ¼ 0:4 and Dx = 0.1 m. The depth gra-
dient method was used for reconstruction of the solution, along with a
wetting-drying procedure to track the position of the shoreline. To simulate the bed
roughness effects during both the run-up and the drawdown, a Manning’s coeffi-
cient n = 0.01 m1/3/s was used, assuming wide channel conditions, the scheme is
implemented in the file solitarywaverunup.xls in Chap. 12. Simulations at different
normalized times T = t(g/ho)1/2 are considered in Fig. 9.29, where the experimental
data of Synolakis (1986) at identical times are also included. The SWE produce an
excellent representation of the wave run-up and drawdown features, including the
formation of the hydraulic jump.
A test for the breaking solitary wave run-up with H/ho = 0.3 is presented in
Fig. 9.30, where the predictions of the SWE are compared with experimental data
of Synolakis (1986). The numerical solution disagrees with the experimental data at
T = 10 and 15, given that non-hydrostatic effects are significant, not accounted for
by the SWE. Wave breaking is observed at T = 20 and 25. Once the wave is
broken, the SWE satisfactorily reproduce the run-up characteristics. At T = 30, 35,
and 40, the SWE predictions are in good agreement with the experimental data
during the run-up process. From T = 45 to 60, a moving hydraulic jump is pro-
gressively formed, linked to the drawdown process. Overall, the bore propagation is
satisfactorily predicted by the SWE model. Further research on this topic was
presented by Hafsteinsson et al. (2017). A sequence of photographs during a model
test is shown in Fig. 9.31, where the run-up flow front and the formation of a
moving hydraulic jump during the drawdown are observed.
As highlighted in this chapter, the finite volume solutions of the 1D SWE
produce results in conformity with exact and experimental results for a variety of
flows, including cases with subcritical, transcritical, and supercritical conditions,
the existence of wet–dry interfaces, uneven topography, steady conditions, and
propagation of bores.
9.9 Computational Examples 411
Fig. 9.30 Breaking solitary wave run-up on a plane beach for H/ho = 0.3 at various normalized
times T = t(g/ho)1/2
412 9 Finite Volume Methods
Fig. 9.31 Model tests of non-breaking solitary wave run-up on slope (VAW photographs)
References
Aureli, F., Maranzoni, A., Mignosa, P., & Ziveri, C. (2008). A weighted surface-depth gradient
method for the numerical integration of the 2D shallow water equations with topography.
Advances in Water Resources, 31(7), 962–974.
Bermudez, A., & Vazquez-Cendón, M. E. (1994). Upwind methods for hyperbolic conservation
laws with source terms. Computers & Fluids, 23(8), 1049–1071.
Bradford, S. F., & Sanders, B. F. (2002). Finite-volume model for shallow-water flooding of
arbitrary topography. Journal of Hydraulic Engineering, 128(3), 289–298.
Bradford, S. F., & Sanders, B. F. (2005). Performance of high-resolution, nonlevel bed,
shallow-water models. Journal of Engineering Mechanics, 131(10), 1073–1081.
Brocchini, M., & Dodd, N. (2008). Nonlinear shallow water equation modeling for coastal
engineering. Journal of Waterway, Port, Coastal, and Ocean Engineering, 134(2), 104–120.
References 413
Brufau, P., Vázquez‐Cendón, M. E., & García‐Navarro, P. (2002). A numerical model for the
flooding and drying of irregular domains. International Journal for Numerical Methods in
Fluids, 39(3), 247–275.
Castro-Orgaz, O., & Chanson, H. (2016). Minimum specific energy and transcritical flow in
unsteady open channel flow. Journal of Irrigation and Drainage Engineering, 142(1),
04015030.
Castro-Orgaz, O., & Chanson, H. (2017). Ritter’s dry-bed dam-break flows: Positive and negative
wave dynamics. Environmental Fluid Mechanics, 17(4), 665–694.
Chaudhry, M. H. (2008). Open-channel flow (2nd ed.). New York: Springer.
Colella, P., & Woodward, P. (1984). The piecewise-parabolic method (PPM) for gas dynamical
simulations. Journal of Computational Physics, 54(1), 174–201.
Cunge, J. A. (1975). Rapidly varying flow in power and pumping canals. In K. Mahmood & V.
Yevjevich (Eds.), Unsteady flow in open channels, 14, 539–586. Fort Collins, CO, USA: Water
Resources Publications.
Cunge, J. A., Holly, F. M., & Verwey, A. (1980). Practical aspects of computational river
hydraulics. London: Pitman.
Dressler, R. (1954). Comparison of theories and experiments for the hydraulic dam-break wave. In
Proceedings of International of Association of Scientific Hydrology, 3(38), 319–328. Rome,
Italy: Assemblée Générale.
Favre, H. (1935). Etude théorique et expérimentale des ondes de translation dans les canaux
découverts [Theoretical and experimental study of travelling surges in open channels]. Dunod,
Paris, France (in French).
Gharangik, A. M., & Chaudhry, M. H. (1991). Numerical simulation of hydraulic jump. Journal of
Hydraulic Engineering, 117(9), 1195–1211.
Glaister, P. (1987). Difference schemes for the shallow water equations (Numerical Analysis
Report 9/87). UK: Department of Mathematics, University of Reading.
Godunov, S. K. (1959). A difference method for numerical calculation of discontinuous solutions
of the equations of hydrodynamics. Matematicheskii Sbornik, 47(3), 271–306 (in Russian).
Gottlieb, S., & Shu, C. W. (1998). Total variation diminishing Runge-Kutta schemes. Mathematics
of Computation, 67(221), 73–85.
Hafsteinsson, H. J., Evers, F. M., & Hager, W. H. (2017). Solitary wave run-up: wave breaking
and bore propagation. Journal of Hydraulic Research, 55(6), 787–798.
Hancock, S. L. (1980). PISCES industrial simulation code manual. Physics International.
Harten, A. (1983). High resolution schemes for hyperbolic conservation laws. Journal of
Computational Physics, 49(3), 357–393.
Harten, A., Lax, P., & van Leer, B. (1983). On upstream differencing and Godunov-type scheme
for hyperbolic conservation laws. Society for Industrial and Applied Mathematics, 25(1), 35–
61.
Henderson, F. M. (1966). Open channel flow. New York: MacMillan Co.
Hirsch, C. (1988). Numerical computation of internal and external flows, 1: Fundamentals of
numerical discretization. Chichester, England: Wiley.
Hirsch, C. (1990). Numerical computation of internal and external flows, 2: Computational
methods for inviscid and viscous flows. Chichester, England: Wiley.
Hoffman, J. D. (2001). Numerical methods for engineers and scientists (2nd ed.). New York:
Marcel Dekker.
Jain, S. C. (2001). Open channel flow. New York: Wiley.
Karni, S. (2011). Nonlinear hyperbolic conservation laws-a brief informal introduction. Lecture
notes on numerical methods for hyperbolic equations: short book course. London: Taylor and
Francis.
414 9 Finite Volume Methods
Khan, A. A., & Lai, W. (2014). Modeling shallow water flows using the discontinuous Galerkin
method. Taylor and Francis, New York: CRC Press.
Lauber, G. (1997). Experimente zur Talsperrenbruchwelle im glatten geneigten Rechteckkanal
[Experiments to the dam break wave in smooth sloping rectangular channel]. Ph.D. thesis, ETH
Zurich, Zürich, Switzerland (in German).
Lax, P. (1954). Weak solutions of nonlinear hyperbolic equations and their numerical
computation. Communications on Pure and Applied Mathematics, 7, 159–193.
Lax, P. D., & Wendroff, B. (1960). Systems of conservation laws. Communications in Pure and
Applied Mathematics, 13(2), 217–237.
LeVeque, R. J. (1992). Numerical methods for conservation laws. Basel, Switzerland: Birkhäuser.
LeVeque, R. J. (2002). Finite volume methods for hyperbolic problems. New York: Cambridge
University Press.
Mingham, C. G., & Causon, D. M. (1998). High-resolution finite-volume method for shallow
water flows. Journal of Hydraulic Engineering, 124(6), 605–614.
Nujic, M. (1995). Efficient implementation of non-oscillatory schemes for the computation of
free-surface flows. Journal of Hydraulic Research, 33(1), 100–111.
Ozmen-Cagatay, H., & Kocaman, S. (2010). Dam-break flows during initial stage using SWE and
RANS approaches. Journal of Hydraulic Research, 48(5), 603–611.
Ozmen-Cagatay, H., & Kocaman, S. (2011). Dam-break flow in the presence of obstacle:
Experiment and CFD simulation. Engineering Applications of Computational Fluid
Mechanics, 5(4), 541–552.
Ritter, A. (1892). Die Fortpflanzung von Wasserwellen [Propagation of water waves]. Zeitschrift
Verein Deutscher Ingenieure, 36(2), 947–954 (in German).
Roache, P. J. (1972). Computational fluid dynamics. Albuquerque NM: Hermosa Publishers.
Roe, P. L. (1981). Approximate Riemann solvers, parameter vectors, and difference schemes.
Journal of Computational Physics, 43(2), 357–372.
Rouse, H. (1961). Fluid motion in a gravitational field. IIHR film, The University of Iowa, Iowa.
Sanders, B. F. (2001). High-resolution and non-oscillatory solution of the St. Venant equations in
non-rectangular and non-prismatic channels. Journal of Hydraulic Research, 39(3), 321–330.
Schoklitsch, A. (1917). Über Dammbruchwellen [On dam break waves]. Kaiserliche Akademie
der Wissenschaften, Wien, Mathematisch-Naturwissenschaftliche Klasse. Sitzungberichte IIa,
126, 1489–1514 (in German).
Sivakumaran, N. S., Tingsanchali, T., & Hosking, R. J. (1983). Steady shallow flow over curved
beds. Journal of Fluid Mechanics, 128, 469–487.
Stoker, J. J. (1957). Water waves: The mathematical theory with applications. New York:
Interscience Publishers.
Sweby, P. K. (1984). High resolution schemes using flux limiters for hyperbolic conservation
laws. Journal of the Society for Industrial and Applied Mathematics, Series B Numerical
Analysis, 21(5), 995–1011.
Sweby, P. K. (1999). Godunov methods (Numerical Analysis Report 7/99), UK: Department of
Mathematics, University of Reading.
Synolakis, C. E. (1986). The runup of long waves (Ph.D. thesis). California Institute of
Technology, Califormia.
Terzidis, G., & Strelkoff, T. (1970). Computation of open channel surges and shocks. Journal of
the Hydraulics Division, ASCE, 96(HY12), 2581–2610.
Toro, E. F. (1992). Riemann problems and the WAF method for solving the two-dimensional
shallow water equations. Philosophical Transactions of the Royal Society of London, Series A,
338, 43–68.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. Singapore: Wiley.
References 415
Toro, E. F. (2009). Riemann solvers and numerical methods for fluid dynamics: A practical
introduction. Berlin, Germany: Springer.
van Albada, G. D., van Leer, B., & Roberts, W. W. (1982). A comparative study of computational
methods in cosmic gas dynamics. Astronomy & Astrophysics, 108(1), 76–84.
van Leer, B. (1979). Towards the ultimate conservative difference scheme, V: A second-order
sequel to Godunov´s method. Journal of Computational Physics, 32, 101–136.
van Leer, B. (2006). Upwind and high-resolution methods for compressible flow: From Donor cell
to residual distribution schemes. Computer Physics Communications, 1(2), 192–206.
Vazquez-Cendón, M. E. (2015). Solving hyperbolic equations with finite volume methods. New
York: Springer.
White, F. M. (2009). Fluid mechanics. New York: McGraw-Hill.
Ying, X., Khan, A., & Wang, S. (2004). Upwind conservative scheme for the Saint Venant
equations. Journal of Hydraulic Engineering, 130(10), 977–987.
Ying, X., & Wang, S. (2008). Improved implementation of the HLL approximate Riemann solver
for one-dimensional open channel flows. Journal of Hydraulic Research, 46(1), 21–34.
Zhou, J. G., Causon, D. M., Mingham, C. G., & Ingram, D. M. (2001). The surface gradient
method for the treatment of source terms in the shallow water equations. Journal of
Computational Physics, 168(1), 1–25.
Zoppou, C., & Roberts, S. (2003). Explicit schemes for dam-break simulations. Journal of
Hydraulic Engineering, 129(1), 11–34.
Chapter 10
Sediment Transport and Movable Beds
10.1 Introduction
Consider 1D unsteady free surface flow over an erodible bed in a vertical plane
(Fig. 10.1). The elevation of the static erodible sediment bed is zb(x, t), and the fluid
flow above is composed of a mixture of water and sediment. The dynamic flow
above the bed is composed of a bed-load layer, whose flux per unit width is qb, and
a suspended-load layer. The latter flow depth is h(x, t), the discharge per unit width
@ @ @Ai
ðqAÞ þ ðqQÞ þ qi ¼ 0; ð10:1Þ
@t @x @t
2
@ @ Q @zs 1 @q
ðqQÞ þ q ¼ qgA qgASf g hp A: ð10:2Þ
@t @x A @x 2 @x
@ @ @zi @zb
ðqhÞ þ ðqhUÞ ¼ qi qb ; ð10:3Þ
@t @x @t @t
@ @ @zs 1 2 @q
ðqhUÞ þ qhU 2 ¼ qgh gh sb : ð10:4Þ
@t @x @x 2 @x
10.2 Flow Equations 419
The bed shear stress is computed with n = Manning’s roughness coefficient and
Rh = hydraulic radius from
n2 U jU j
sb ¼ qg 1=3
: ð10:5Þ
Rh
@ @
ðhCs Þ þ ðhUCs Þ ¼ E D: ð10:6Þ
@t @x
@ @qb @zb
ðhb Cb Þ þ þ ð1 pm Þ ¼ D E; ð10:7Þ
@t @x @t
@zb 1 qb qb
¼ DEþ ; ð10:8Þ
@t 1 pm L
@ @qb qb qb
ðhb Cb Þ þ ¼ : ð10:9Þ
@t @x L
420 10 Sediment Transport and Movable Beds
@h @ 1 qb qb
þ ðhUÞ ¼ E Dþ ; ð10:10Þ
@t @x 1 pm L
@ @ 2 @zs sb 1 g 2 @q
ðhUÞ þ hU ¼ gh h
@t @x @x q 2 q @x
ð10:11Þ
q q U qb qb
b E Dþ ;
q 1 pm L
Rather than using Eqs. (10.6)–(10.7), it is possible to define a transport equation for
the total load if separate computations for each load are of no interest. Summing
Eqs. (10.6)–(10.7) produces
@ @ @zb
ðhCs þ hb Cb Þ þ ðhUCs þ qb Þ þ ð1 pm Þ ¼ 0: ð10:12Þ
@t @x @t
@ @ qb qb
ðhCt Þ þ ðhUCt Þ ¼ D E þ : ð10:13Þ
@t @x L
We now define the source term of Eq. (10.13) as function of the total load as (Wu
2008)
10.2 Flow Equations 421
qb qb qt qt
DEþ w¼ ; ð10:14Þ
L L
where qt* = total-load sediment transport capacity rate, qt = qCt = actual total-load
sediment transport rate, and L = non-capacity adaptation length of the total-load
sediment transport.
The settling velocity of sediment particles in turbid water xs is determined from the
Richardson–Zaki formula with m = 4 (Wu 2008)
xs ¼ xo ð1 Ct Þm ; ð10:17Þ
where the settling velocity of a single particle in clear water is with d as the particle
diameter and m the kinematic viscosity of water (Wu 2008)
h i0:5
xo ¼ ð13:95m=d Þ2 þ 1:09½ðqs =qÞ 1gd 13:95m=d: ð10:18Þ
The actual total load is qt= hUCt, while the total-load capacity is determined as
the contribution of both the suspended-load and the bed-load as
Here, qs* and qb* = suspended- and bed-load sediment transport capacity rates,
respectively, computed using Wu et al.’s (2000) formulas. To compute the equi-
librium bed-load flux qb*, alternative empirical formulations could be considered,
including the classical equations by Meyer-Peter and Müller, Yalin and van Rijn
(Dey 2014).
422 10 Sediment Transport and Movable Beds
where d = single sediment size diameter, s = shear stress on the wetted perimeter
of the cross-section, sb = bed shear stress tangent to sediment–fluid interface,
sc = critical bed shear stress determining the threshold for incipient sediment
motion, n′ = Manning’s coefficient corresponding to the grain roughness for a
movable bed (= d1/6/20), and n = Manning’s roughness coefficient for the channel
bed, with d in [m] and n in [sm−1/3]. Assuming a channel width larger than the flow
depth, s sb. By numerical experimentation in shallow water flows, it was found
that the last assumption does not reduce the accuracy of the solution. The critical
shear stress is calculated with H = Shields parameter = 0.03 as (Wu 2008)
sb qu2
¼ 2 : ð10:23Þ
s c q u c
g1=2 U
u ¼ n 1=6
; ð10:24Þ
Rh
Note that K is based on the depth-averaged mixture density. Fraccarollo and Capart
(2002) observed extremely high sediment concentration near the bed in geomorphic
dam break flows. Thus, for dam break waves, the empirically enhanced version of
K used is
424 10 Sediment Transport and Movable Beds
qb
K ¼ ½ð1 Ct Þ þ sCt b : ð10:27Þ
qw
The near-bed value for dam break waves over erodible beds considered below is
Ct = 0.6.
Equations (10.10), (10.11), (10.13), and (10.8) can be written as (Cao et al. 2004;
Wu and Wang 2008)
@U @F
þ ¼ S; ð10:28Þ
@t @x
where
0 w
1
1p
0 1 0 1 B m
@zs sb 1 g 2 @q C
B C
h hU B gh h C
B hU C B hU 2 C B @x q 2 q @x C
B C
U¼B C
@ hCt A; F¼B C
@ hUCt A; S¼B q q U C:
B þ b w C
B q 1 pm C
zb 0 B C
@ w A
w
1pm
ð10:29Þ
Here, U is the dependent variable vector, F the flux in the x-direction, and S the
source term. Note that the thickness of the bed-load layer is not resolved by this
model, as originally assumed, that is, zs = zb+ h.
Consider a rectangular finite volume mesh in the x-t plane. Cell i is limited by both
cell interfaces i − 1/2 and i + 1/2. In the t-axis, the control volumes are delimited
by the time levels k and k + 1. Integration of Eq. (10.28) over a control volume
yields exactly (Toro 2001) [see Eq. (9.7)]
10.3 Numerical Scheme 425
Dt
Uki þ 1 ¼ Uki Fi þ 1=2 Fi1=2 þ DtSi ; ð10:30Þ
Dx
where k = time level, Δt = time step, and Δx = grid step. According to Ying et al.
(2004), the one-sided upwind method is used to evaluate the intercell fluxes in
Eq. (10.30) as
0 1k
h
B hU 2 C
Fi þ 1=2 ¼B C
@ hUCt A : ð10:31Þ
0 iþl
To increase model robustness, the flow depth in the momentum topographic source
term is evaluated at the time level k + 1, given that it is available by previously
solving the continuity equation. To ensure numerical stability of the one-sided
first-order upwind method, the water surface gradient in Eq. (10.30) is evaluated
following Ying et al. (2004) by using a weighted average of the downwind and the
upwind gradients as
!
@zs zks;iþþ11 zks;iþ 1 zks;iþ 1 zks;i1
þ1
gh ghki þ 1 w1 þ w2 : ð10:32Þ
@x i Dx Dx
max
Ui þ ðghi Þ
The maximum value for CFL was determined by trial-and-error numerical exper-
iments. The value was typically less than 0.5.
Dam break flow is a topic of continued research interest given its highly detrimental
effects. The usual engineering approach to predict these flows relies on the use of
10.4 Test Cases 427
Fig. 10.3 Water and bed surfaces at different times for the Taipei test case
In Fig. 10.3, the Taipei experiment (Capart and Young 1998) is simulated by
solving Eqs. (10.28)–(10.29) at times t = 3to, 4to, and 5to after the dam failure,
where the time scaling is to = (ho/g)1/2 0.1 s. The sediment particles used were
artificial pearls covered with a shiny white coating, having d = 6.1 mm, qs =
1048 kg/m3, and xo = 7.6 cm/s. The flume was sufficiently long and deep with
b = 0.2 m. The model produces a bed profile eroded by the dam break flow in
overall agreement with the experiments. The experimental free surface profile is
composed of a negative smooth wave, followed by a train of undulations above the
10.4 Test Cases 429
scour hole, ending in a positive wave where its edge is a wet–dry front propagating
in the positive x-direction. Note that the experimental data show free surface
undulations at the vicinity of the positive wave portion, not predicted by hydrostatic
computations. However, the position of the wet–dry front predicted by the
hydrostatic model is in agreement with the experimental data. The overall wave
features are thus well predicted by the SWE.
The study of the Louvain test experiment at times t = 5to, 7.5to, and 10to after
the dam break is presented in Fig. 10.4. In this experiment, Fraccarollo and Capart
Fig. 10.4 Water and bed surfaces at different times for Louvain test case
430 10 Sediment Transport and Movable Beds
(2002) used cylindrical PVC pellets as sediment particles, having d = 3.5 mm,
qs = 1540 kg/m3, and xo = 18 cm/s. The flume had a width b = 0.10 m. The
hydrostatic model produces simulations of the eroded bed profiles with some
divergence from the experimental results, but in overall agreement. The free surface
profile predicted is in fair agreement with observations, but the undulations due to
non-hydrostatic pressure are not reproduced (Cantero-Chinchilla et al. 2016).
Dike breach experiments of Schmocker (2011) were selected for comparison purposes
with numerical simulations using Eq. (10.30). Figure 10.5 shows a definition sketch of
his experimental setup. Tests 42 and 55 were selected, involving variations in sediment
diameter (non-cohesive material) and discharge. The tests features are summarized in
Table 10.1. The dike erosion process produces a round-crested weir-like profile
(Fig. 10.6), with features different from the dam break wave problem.
The experimental data of Test 42 are compared in Fig. 10.7 with the numerical
solution resulting from Eq. (10.30). Values of Dx = 0.025 m and CFL ¼ 0:25
were set in the numerical model, which is implemented in the file
Movablebed_Dikebreaching.xls, available in Chap. 12. For the simulations K = 1
[see Eq. (10.26)], L = 0.01 m, xs = xo, and n = 0.018 s/m1/3 were used. The bed
sediment layer porosity pm is calculated from (Wu and Wang 2008)
Fig. 10.5 Definition sketch of experimental setup (Schmocker 2011), subscript “0” defines the
inflow conditions
Fig. 10.7 Water and bed surfaces at different times for dike erosion Test 42
432 10 Sediment Transport and Movable Beds
0:21
pm ¼ 0:13 þ : ð10:36Þ
½d ðmmÞ þ 0:0020:21
The dry bed was simulated by adopting the minute flow depth 0.0001 m. The
upstream discharge was implemented in the numerical model as an instantaneous
pulse in a ghost cell, whereas the upstream water depth at the ghost cell was
determined by extrapolating the interior solution at the previous instant. At the
downstream section, transmissive boundary conditions were implemented in a
ghost cell (Toro 2001). For Test 42, the discharge is low, and non-hydrostatic
effects are expected to be small. This is in turn verified by comparing simulations
with measurements in Fig. 10.7, from where it is observed that the water level
upstream of the dike crest is in conformity with the theory. Beyond the dike crest,
the numerical model predicts excessive erosion on the downstream slope, thereby
diverging theory from experimental results.
Fig. 10.8 Water and bed surfaces at different times for dike erosion Test 55
10.4 Test Cases 433
References
Cantero-Chinchilla, F., Castro-Orgaz, O., Dey, S., & Ayuso, J. L. (2016). Nonhydrostatic dam
break flows II: One-dimensional depth-averaged modelling for movable bed flows. Journal of
Hydraulic Engineering, 142(12), 04016069.
Cantero-Chinchilla, F. N., Castro-Orgaz, O., Schmocker, L., Hager, W. H., & Dey, S. (2018).
Depth-averaged modelling of granular dike overtopping. Journal of Hydraulic Research,
56(4), 537–550.
Cantero-Chinchilla, F. N., Castro-Orgaz, O., Dey, S. (2019). Prediction of overtopping dike
failure: A sediment transport and dynamic granular bed deformation model. Journal of
Hydraulic Engineering, 145(6), 04019021.
Cao, Z., Pender, G., Wallis, S., & Carling, P. (2004). Computational dam-break hydraulics over
erodible sediment bed. Journal of Hydraulic Engineering, 130(7), 689–703.
Capart, H., & Young, D. L. (1998). Formation of a jump by the dam-break wave over a granular
bed. Journal of Fluid Mechanics, 372, 165–187.
Capart, H., Young, D. L. (2002). Two-layer shallow water computations of torrential flows.
In: Proceedings of River Flow, vol. 2, (pp. 1003–1012). Lisse, The Netherlands: Balkema.
Dey, S. (2014). Fluvial hydrodynamics: Hydrodynamic and sediment transport phenomena.
Berlin: Springer.
Fraccarollo, L., & Capart, H. (2002). Riemann wave description of erosional dam break flows.
Journal of Fluid Mechanics, 461, 183–228.
Greco, M., Iervolino, M., Leopardi, A., & Vacca, A. (2012). A two-phase model for fast
geomorphic shallow flows. International Journal of Sediment Research, 27(4), 409–425.
Pontillo, M., Schmocker, L., Greco, M., & Hager, W. H. (2010). 1D numerical evaluation of dike
erosion due to overtopping. Journal of Hydraulic Research, 48(5), 573–582.
Schmocker, L. (2011). Hydraulics of dike breaching. Ph.D. thesis. Zürich, Switzerland: Swiss
Federal Institute of Technology.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. New York: Wiley.
Wu, W. (2008). Computational river dynamics. London, U.K.: Taylor and Francis.
Wu, W., & Wang, S. S. Y. (1999). Movable bed roughness in alluvial rivers. Journal of Hydraulic
Engineering, 125(12), 1309–1312.
434 10 Sediment Transport and Movable Beds
Wu, W., & Wang, S. S. Y. (2007). One-dimensional modeling of dam-break flow over movable
beds. Journal of Hydraulic Engineering, 133(1), 48–58.
Wu, W., & Wang, S. S. Y. (2008). One-dimensional explicit finite-volume model for sediment
transport. Journal of Hydraulic Research, 46(1), 87–98.
Wu, W., Wang, S. S. Y., & Jia, Y. (2000). Nonuniform sediment transport in alluvial rivers.
Journal of Hydraulic Research, 38(6), 427–434.
Wu, W., Vieira, D. A., & Wang, S. S. Y. (2004). One-dimensional numerical model for
nonuniform sediment transport under unsteady flows in channel networks. Journal of
Hydraulic Engineering, 130(9), 914–923.
Ying, X., Khan, A. A., & Wang, S. S. Y. (2004). Upwind conservative scheme for the Saint
Venant equations. Journal of Hydraulic Engineering, 130(10), 977–987.
Chapter 11
Numerical Modeling of Non-hydrostatic
Free Surface Flows
11.1 Introduction
The shallow water equations (SWE) are a dispersionless system of hyperbolic PDEs
obtained by assuming that the vertical flow acceleration is negligible. Under this
assumption, the vertical momentum balance is reduced to the hydrostatic pressure
law (Toro 2001; Castro-Orgaz and Hager 2017). This approach produces good
solutions for shallow flows if the vertical length scale [H] is negligible as compared
to the horizontal length scale [L]. In turn, this scenario is realistic in many open
channel and river hydraulics problems. However, there are as well a large number
of practical questions where this level of mathematical approximation is not well
suited. It includes flows across hydraulic structures like weirs, intakes, overfalls,
energy dissipators, and water wave motion in maritime and fluvial hydraulics, as the
solitary and cnoidal waves, and the undular bore propagation in a river. In this large
portfolio of problems, the vertical length scale [H], though it can be small, is not
negligible as compared to [L] (Steffler and Jin 1993; Castro-Orgaz and Hager 2017).
A non-hydrostatic vertical pressure distribution must therefore be accounted for in
the depth-averaged equations. A higher-order mathematical model based on the
so-called Boussinesq equations is a possible choice to simulate these flows.
Depending on the technique and terms retained while introducing the vertical
acceleration and/or turbulent stresses on the depth-averaged equations, different
types of Boussinesq equations are obtained. In this chapter, we only consider
inviscid flow solutions following Castro-Orgaz and Cantero-Chinchilla (2019). The
non-hydrostatic flows considered are shallow, and waves are thus long and dis-
persion effects weak. The modeling of highly dispersive waves from deep to
intermediate depths is beyond the scope of this chapter. First, steady potential open
channel flow problems in a vertical plane are considered, then generalizations for
inviscid unsteady flow over a 3D terrain. Illustrative problems relating to hydraulic
structures and wave motion are solved using second- and higher-order numerical
schemes.
Consider steady two-dimensional (2D) potential free surface flows over a curved
bed (Fig. 11.1). The governing equations are (Rouse 1938; Vallentine 1969)
@u @w
þ ¼ 0; ð11:1Þ
@x @z
@u @u 1 @p
u þw ¼ ; ð11:2Þ
@x @z q @x
@w @w 1 @p
u þw ¼ g; ð11:3Þ
@x @z q @z
@u @w
¼ 0: ð11:4Þ
@z @x
@/ @/
u¼ ; w¼ : ð11:5Þ
@x @z
This flow field automatically satisfies Eq. (11.4). Inserting Eq. (11.5) in Eq. (11.1),
the flow field obeys the Laplace equation for /,
@2/ @2/
þ 2 ¼ 0: ð11:6Þ
@x2 @z
@w @w
u¼ ; w¼ : ð11:7Þ
@z @x
Using these relations in Eq. (11.1), the continuity equation is automatically satis-
fied. Using Eq. (11.7) in Eq. (11.4), the stream function satisfies the Laplace
equation
@2w @2w
þ 2 ¼ 0: ð11:8Þ
@x2 @z
The lines w = const. are streamlines, and the velocity vector, of modulus
V = (u2 + w2)1/2 and inclined with respect to the x-axis by the angle h = tan−1(w/u),
is tangent to them. The equipotential lines / = const. are normal to the streamlines
at intersection points, forming the so-called flow net of a potential flow (Fig. 11.1).
The Cauchy–Riemann equations are derived by equaling the velocity components
in Eqs. (11.5) and (11.7) as
@/ @w @/ @w
u¼ ¼ ; w¼ ¼ : ð11:9Þ
@x @z @z @x
A final result of relevance to solve potential free surface flows is the Bernoulli
equation, stating conservation of the total energy head H within the fluid domain
(Vallentine 1969)
p u2 þ w2
H ¼ zþ þ ¼ const: ð11:10Þ
qg 2g
@zs @zb
ws ¼ us ; w b ¼ ub : ð11:11Þ
@x @x
dS pb dzb
¼ ; ð11:12Þ
dx qg dx
where
b þh
zZ
u2 p
S¼ þ dz ð11:13Þ
g qg
zb
with the variable η(x, z) = z − zb(x) as the vertical distance above the channel
bottom, U = q/h the depth-averaged velocity, h the vertical flow depth, and q the
unit discharge. Subscripts indicate differentiation, e.g., Ux =@U=@x, Uxx =@ 2 U=@x2 ,
ηx = −@zb =@x, and ηxx = −@ 2 zb =@x2 . For steady flow, the spatial derivatives of
U and η take the form
Inserting Eq. (11.16) into Eqs. (11.14) and (11.15) results in (Castro-Orgaz and
Hager 2017)
2
q 2hx zbx 2g h hxx h2x 3g h2
u ¼ 1 þ zbxx þ ; ð11:17Þ
h h 2 2h h2 3
qh g i
w ¼ zbx þ hx : ð11:18Þ
h h
This theory considers the curvature effect of both the free surface and the bottom,
accounted for by the inclusion of hxx = d2h/dx2 and zbxx = d2zb/dx2; it also considers
the slope effect of the free surface and the bottom by inclusion of hx = dh/dx and
zbx = dzb/dx. The values of these slopes and curvatures are not necessarily small,
given that no restriction on non-linearity was imposed. Therefore, Matthew’s
(1991) theory, used by Castro-Orgaz and Hager (2009), is an approximate potential
flow model that includes the effects of finite curvatures and slopes. This equation
was presented by Naghdi and Vongsarnpigoon (1986) based on the theory of a
Cosserat surface (Green and Naghdi 1976a, b; Naghdi 1979). It is remarkable to
state that this equation is the steady-state form of the well-known Serre-Green-
Naghdi equations, which are widely applied in ocean research. This issue will be
further exploited below. Equation (11.19) was also obtained by Marchi (1992,
1993) by expanding the stream function in power series, and by Zhu and Lawrence
(1998) by using a perturbation method. It is a second-order differential equation,
from which the free surface profile h = h(x) is determined. For given q and
H = const., and prescribed flow depths at two boundary sections, Eq. (11.19) is
440 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
pb q2
¼ hþ 2hzbxx þ hhxx h2x 2zbx hx : ð11:21Þ
c 2gh 2
After having established the energy equation, the momentum principle is con-
sidered. Inserting Eqs. (11.17) and (11.20) for the distributions of u and p into
Eq. (11.13), the specific momentum S is given by (Castro-Orgaz and Hager 2017)
[see Eq. (1.133)]
h2 q2 hhxx h2x hzbxx hx zbx
S¼ þ 1þ þ : ð11:22Þ
2 gh 3 2 2
Using Eqs. (11.21) and (11.22), the vertically integrated momentum balance
[Eq. (11.12)] can be equally used to compute the free surface profile h = h(x). It
remains to delineate the connection between the energy [Eq. (11.19)] and
momentum [Eqs. (11.12), (11.21) and (11.22)] models for steady potential flow
over curved beds. Consider first Eq. (11.19), given that the energy head is a con-
stant in potential flow, its differentiation with respect to x yields
dH q2 2hhxx h2x
¼ zbx þ hx 3 hx 1 þ þ hzbxx þ z2bx
dx gh 3
ð11:23Þ
q2 2hhxxx
þ þ hz bxxx þ h z
x bxx þ 2z z
bx bxx ¼ 0:
2gh2 3
q2 2 q2 1 q2 3 1 q2
zbx þ hx h x h x h xx þ h þ hxxx
gh3 3 gh2 3 gh3 x 3 gh
ð11:24Þ
1 q2 q2 1 q2 q2
h x z bxx h x z2
þ z bxxx þ zbx zbxx ¼ 0:
2 gh2 gh3 bx 2 gh gh2
d h2 q2 hhxx h2x hzbxx hx zbx
þ 1þ þ
dx 2 gh 3 2 2
q2 dzb
¼ hþ 2hzbxx þ hh xx h 2
x 2z h
bx x : ð11:25Þ
2gh2 dx
Expanding the various terms in Eq. (11.25) yields Eq. (11.24). This implies that the
energy and momentum balances produce the identical governing equation to this
order of accuracy. Therefore, it is in principle easier to adopt Eq. (11.19), which is
an integral form of Eq. (11.25). This is interesting from a numerical standpoint,
given that the higher order of differentiation is reduced to second order, avoiding
instability problems linked to the discretization of third-order derivatives in
Boussinesq-type equations (Bonneton et al. 2011). Equation (11.25) is the
steady-state version of the Serre–Green–Naghdi equations; thus, Eq. (11.19) is an
integral form of these equations under steady flow. Equation (11.19) admits ana-
lytical solutions for selected problems like the flow profile upstream of a free
overfall, the solitary wave, or the supercritical jet beyond a sluice gate
(Castro-Orgaz and Hager 2017). However, in general, a numerical solution is
required for flows over curved beds.
This relation is valid at the spillway crest (Fig. 11.2); accepting a physical solution
with finite free surface slope (hx < 0), it is the critical flow condition for curvilinear
flows. This condition serves as the mathematical equation describing the critical
depth in non-hydrostatic flows (Castro-Orgaz and Hager 2017).Equation (11.26)
indicates that the non-hydrostatic critical flow condition does not depend on the
section conditions alone, e.g., the actual value of h, but in addition on the con-
figuration of the flow profile h = h(x) in the vicinity of the crest section, and on the
local bottom geometry variation zb = zb(x). The words by Rouse (1938) on p. 326
of his book state:
… Since the ratio of depth to specific energy at the true critical section is so definitely a
function of the curvature imposed by the fixed boundaries, it is almost futile to expect that a
simple relationship may be found expressing this ratio in terms of boundary geometry. It is
to be hoped, nevertheless, that a broader understanding of true critical discharge may soon
lead to definite progress in this essential field.
442 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Equation (11.26) is an advance in this line; the non-hydrostatic effects imposed by the
boundary conditions are included in the critical flow statement. Equation (11.26) is
an approximate solution at the weir crest to the elliptic problem posed by Laplace’s
equation for potential flow. Therefore, the approximation to this elliptic problem
depends not only on the crest conditions, but rather on the complete flow solution,
which, in turn, determines the spatial derivatives hx, hxx, and hxxx.
It is possible, however, to find approximations to these derivatives at the weir
crest on the basis of the lower-order energy head equation for hydrostatic flows
(Hager 1985; Matthew 1991) [Eq. (2.12)]
q2
H ¼ zb þ h þ ¼ const: ð11:27Þ
2gh2
This theory for critical curvilinear flows is presented here following Castro-Orgaz
and Hager (2017). Differentiation of Eq. (11.27) produces for critical flow (F ¼ 1)
the following approximations for h2x and hhxx at the weir crest, with R as the crest
curvature radius (Castro-Orgaz and Hager 2017),
hzbxx h
h2x ¼ ¼ ; ð11:28Þ
3 3R
4zbxx 4
hxx ¼ ¼ : ð11:29Þ
9 9R
11.2 Two-Dimensional Steady Potential Flow 443
The specific energy at the weir crest (dzb/dx = 0) is from Eq. (11.19)
q2 2hhxx h2x
E ¼ hþ 1þ þ hzbxx ; ð11:30Þ
2gh2 3
Inserting Eqs. (11.28)–(11.29) into Eq. (11.30), and using Eq. (11.31) produce,
retaining first-order terms (Castro-Orgaz and Hager 2017)
3=2
2 22 E
Cd ¼ 1þ : ð11:32Þ
3 81 R
The critical depth in curvilinear motion is determined from Eq. (11.26). The
third-order flow depth derivative term needed in this equation is approximated by
differentiation of Eq. (11.27) as (Castro-Orgaz and Hager 2017)
h2 hxxx 5 5h
¼ hzbxx ¼ : ð11:33Þ
hx 9 9R
Inserting Eqs. (11.28), (11.29), and (11.33) into Eq. (11.26) produces when
retaining first-order terms (Castro-Orgaz and Hager 2017)
hcrest E
¼1 : ð11:34Þ
hc 9R
The bottom pressure head at the weir crest is from Eq. (11.21)
pb q2
¼ hþ 2hzbxx þ hhxx h2x ; ð11:35Þ
c 2gh 2
where c is the specific weight of fluid. Using Eqs. (11.28), (11.29), and (11.34) into
Eq. (11.35) yields to first-order accuracy
Equations (11.32), (11.34), and (11.36) comprise the first-order solutions for the
problem of critical flow in curvilinear motion. They are expected to be valid for
weakly curved flows, e.g., up to E/R = 1 (Castro-Orgaz and Hager 2017). This
444 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
theory is compared in Fig. 11.3 with the experimental data of Blau (1963) for a
parabolic weir of crest radius R = 0.919 m. Note that the approximate theory
produces a solution in fair agreement with observations. The advantage of this type
of model is that the full solution of Eq. (11.19) is avoided. The full solution h(x) of
flow over curved beds can only be obtained numerically, given the lack of known
exact analytical expressions.
11.2 Two-Dimensional Steady Potential Flow 445
Consider flows away from a slope break, where streamline curvature effects can be
neglected. The flow is gradually varied on a steep slope (Fig. 11.4), given that the
variation of h with x is small, as confirmed by experiments (Castro-Orgaz and
Hager 2009, 2017).
For these flows, it can be assumed that h2x hxx 0. Further, on the slope, the
bed is flat, resulting in zbxx = 0. On this slope, however, the term zbx is finite.
Therefore, despite hx will be small, the product (hx ⋅ zbx) remains finite. Therefore,
Eqs. (11.19) and (11.21) for gradually varied, 1D potential flow on a finite slope
read (Castro-Orgaz and Chanson 2016)
q2
H ¼ zb þ h þ 2
1 þ z2bx ; ð11:37Þ
2gh
pb q2
¼h ð2zbx hx Þ: ð11:38Þ
c 2gh2
Differentiation of Eq. (11.37) produces for supercritical flow on the steep slope
(large F) the ODE (Castro-Orgaz and Hager 2009)
dh zbx
¼ q2
: ð11:39Þ
dx
gh3 1 þ z2bx
Inserting Eq. (11.39) into Eq. (11.38), the bottom pressure head is [see Eq. (1.137)]
pb q2 z2bx h
¼h ð 2z h
bx x Þ ¼ h 1 ¼ : ð11:40Þ
c 2gh 2 1 þ zbx2 1 þ z2bx
Equation (11.40) is the bottom pressure head in gradually varied flows on a steep
slope, implying non-hydrostatic conditions. This is the value to which 1D com-
putations tend on a slope using Eq. (11.19) (Castro-Orgaz and Hager 2009).
On inspecting Eqs. (11.36) and (11.40), it can be appreciated that Matthew’s (1991)
theory accounts for curvature (weir crest) and slope (chute) effects in non-
hydrostatic fluid pressure. The theory has the advantage that it is mathematically
elaborated. Thus, the salient results are rigorous statements, without the need of
artificial tuning of the final equations, based on ad hoc assumptions. A numerical
method to solve Eq. (11.19) is considered in this section for the flow over a
round-crested weir. The method will be presented taking as test case transcritical
flow over a bell-shaped hump prescribed by
x 2
zb ¼ a exp b : ð11:41Þ
L
Fig. 11.5 Finite-difference model for the solution of the steady Serre–Green–Naghdi equations
[adapted from Castro-Orgaz and Cantero-Chinchilla (2019)]
11.2 Two-Dimensional Steady Potential Flow 447
solution. The nodes i = 1 and i = N are located away from the hump in zones where the
streamlines are parallel and the pressure is hydrostatic. Therefore, H is estimated from
q2
H h1 þ : ð11:42Þ
2gh21
However, neither h(i = 1) nor h(i = N) is known in advance for weir flow. For a
transition from upstream subcritical to downstream supercritical flow, these flow
depths are the alternate depths corresponding to the total head H. For a numerical
solution, an iterative strategy is necessary. Equation (11.19) can be solved numeri-
cally using a shooting method transforming it into a pair of ODEs. Starting with a
guessed value of h1 and setting dh/dx = 0 there, given that the streamlines are parallel
at the inflow section (Castro-Orgaz and Hager 2017), the flow profile is explicitly
computed based on the results at the previous computational section. Once h1 is
assumed, the corresponding value of H is determined by Eq. (11.42). Using a Runge–
Kutta solver, the system of two ODEs is integrated along the computational domain
from i = 1 up to i = N, with the flow depth there determined as the final result of the
numerical integration. This value of the flow depth is compared to the supercritical
alternate depth of h1, and, if they do not match within a prescribed tolerance, then h1
and thus also H are both in error; the value of h1 must be therefore corrected. This
method of attack for transcritical flow over a round-crested weir was first discussed
by Naghdi and Vongsarnpigoon (1986), who numerically solved Eq. (11.19) for the
weir experiments of Sivakumaran et al. (1983). They found that if the value of h1 is
too low, the flow profile intersects the bed profile elsewhere (h = 0 at this point), and
computations are thus aborted before reaching the end section. If the head is too large,
an undular jump is formed at the tailwater portion of the weir. The value of h1 can be
iterated, but a high precision is needed to obtain the correct physical solution. This
was later discussed by Fenton (1996) and Castro-Orgaz and Hager (2017).
Castro-Orgaz and Hager (2013) demonstrated that a physically correct solution for
transcritical weir flow follows with this shooting technique.
An alternative method of solution is to treat Eq. (11.19) as a boundary value
problem directly, discretizing the equations using finite-differences and solving the
implicit system of equations iteratively. The first solution of a steady
Boussinesq-type equation using an implicit finite-difference method was elaborated
by Hosoda and Tada (1994) to compute the undular jump profile. For the specific
case of transcritical flow over a weir, Zhu (1996) and Zhu and Lawrence (1998)
solved Eq. (11.19) as a boundary value problem using a collocation method,
finding excellent agreement between the numerical solution and the data of
Sivakumaran et al. (1983). Onda and Hosoda (2004) solved Eq. (11.15) for flow
over curved bed forms, but added terms to account for bed friction and turbulence.
They used a second-order accurate central finite-difference scheme and solved the
resulting equations iteratively with a technique similar to a Newton–Raphson
method. Zerihun and Fenton (2006) computed transcritical flows with an implicit
finite-difference scheme solved iteratively with a Newton–Raphson iteration
assisted by a lower-upper (LU) decomposition of an analytical Jacobian matrix.
448 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Castro-Orgaz and Hager (2009) applied Eq. (11.19) for the transition from a
mild to a steep slope using a centered finite-difference scheme fourth-order accurate
and solved the resulting implicit system of equations as an optimization problem.
Here, a simpler version of the finite-difference method of Castro-Orgaz and Hager
(2009) is elaborated. Following Onda and Hosoda (2004), the free surface
derivatives of steady Boussinesq-type equations can be discretized using second-
order central finite-differences as
Inserting Eq. (11.43) into Eq. (11.19) yields the algebraic equation at node i as
"
q2 2hi ðhi þ 1 hi1 Þ2
H i ¼ ð z b Þ i þ hi þ 2
1þ 2
ðhi þ 1 2hi þ hi1 Þ
2ghi 3ðDxÞ 12ðDxÞ2
#
þ hi ðzbxx Þi þ z2bx i :
ð11:44Þ
If the discrete flow depths hi−1, hi, and hi+1 are an accurate representation of the true
flow solution at these nodes, then the estimated head at node Hi can be considered
to equal H, within a prescribed tolerance. Thus
Here, e is the tolerance, typically set to 10−6. In this event, the problem is finished
and there is no need to iterate the flow depths at the finite-difference nodes.
However, the initially assumed distribution of flow depths at the nodes of the mesh
is in general not accurate enough, and there is a need of iterative solutions. Thus, at
each node i, the error ei generated by the assumed solution is
ei ¼ H Hi [ eH: ð11:46Þ
Now the correction strategy is elaborated. Let k be the iteration index, then the
variation of the error at node i between two consecutive iterations can be expressed
using a truncated Taylor series as
@ei @ei @ei
dei ¼ eki þ 1 eki ¼ dhi1 þ dhi þ dhi þ 1 : ð11:47Þ
@hi1 @hi @hi þ 1
We seek setting to zero the error after the current iteration, that is, eki þ 1 0. Thus,
Eq. (11.47) for the N − 2 computational nodes is written in matrix form with J as
the Jacobian matrix as
11.2 Two-Dimensional Steady Potential Flow 449
Thanks to the second-order differences used [Eq. (11.43)] to discretize the steady
Boussinesq-type equation [Eq. (11.19)], the structure of the linear system (11.48) is
tridiagonal, and the vector of corrections dh can be efficiently determined using the
Thomas algorithm (Hoffman 2001). Once this task is accomplished, the new esti-
mation of the flow depths is given by
hk þ 1 ¼ hk þ dh: ð11:49Þ
q2
H ¼ zb þ h þ ; ð11:53Þ
2gh2
yields an estimation of flow depths in the subcritical flow portion, which are too far
away from the Boussinesq solution for some tests, however. This produces diver-
gent numerical computations.
A different strategy was therefore adopted based on Eq. (11.53). The assumed
flow depth h1 was used to estimate the head as H = h1; the specific energy at the
weir crest is then Emin = h1 − a. The hydrostatic discharge compatible with this
minimum specific energy is (Chanson 2004)
450 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
3=2
2 3 1=2
qo ¼ gEmin : ð11:54Þ
3
with hc = (q2o/g)1/3. The transcritical hydrostatic flow profile computed for qo was
found to produce a reasonable initial estimate of the flow depths at the
finite-difference nodes. Once the free surface profile is accurately predicted with
Eq. (11.19), the result must be analyzed. If a train of standing waves is formed
upstream from the weir, this points at an incorrect estimate of H. It basically
indicates that the corresponding value of the discharge coefficient Cd is incorrect,
with
q
Cd ¼ h i1=2 : ð11:58Þ
gð H aÞ 3
For transcritical weir flow, the flow passes from upstream subcritical to down-
stream supercritical flow without appreciable waves on the flow profile. The rela-
tion between q and H is unique, and the value of Cd is therefore fixed. However, for
a given q, other flow profiles with upstream waves are mathematically possible
from Eq. (11.19) for different values of H. These are not physical solutions to the
weir flow problem, however. This phenomenon is described by Zhu (1996), who
also found upstream trains of standing waves while solving Eq. (11.19) as a
boundary value problem using a collocation method. The head H (or h1) must be
iterated until the upstream waves are suppressed.
The process of numerical solution is as follows:
1. Given the discharge q, a value of h1 is assumed, and H is then computed using
Eq. (11.42).
2. The value of hN is the supercritical alternate depth of h1.
11.2 Two-Dimensional Steady Potential Flow 451
where E = H − a. For the Gaussian hump, d2zb/dx2 = –2ab/L2 at the crest. If the
assumed head is too far from its correct value, the numerical solution may diverge
while updating h. This was avoided using a relaxation factor of 0.25 in the Newton–
Raphson algorithm.
Figure 11.6 displays the experimental data of Sivakumaran et al. (1983) for a
symmetrical hump of shape zb = 0.2 exp[−0.5(x/0.24)2] (m). The unit discharge is
0.11197 m2/s. The up- and downstream boundary sections are located at
452 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Fig. 11.6 Iteration of upstream flow depth in transcritical non-hydrostatic flow over a weir a RUN
1, b RUN 2 (see Table 11.1) [adapted from Castro-Orgaz and Cantero-Chinchilla (2019)]
x = ±1.5 m. Equation (11.19) was numerically solved with the implicit finite-
difference method described above using Δx = 0.05 m. A first computation was
conducted assuming h1 = 0.35 m, and the corresponding free surface profile and
discharge coefficient are reported in Fig. 11.6a and Table 11.1 as RUN1, respec-
tively. Note that this arbitrary h1 value, though close to experiments, produces an
inacceptable variation of Cd, with the corresponding train of upstream waves
(Fig. 11.6a). For RUN 2, the value of h1 after some iterations is given in
Table 11.1, resulting in a smooth free surface profile (Fig. 11.6b). The predicted Cd
value is now very close to the experimental value.
Once the final free surface profile is determined, the bottom pressure head is
estimated from the finite-difference version of Eq. (11.21) as
"
pb q2 hi
¼ hi þ 2hi ðzbxx Þi þ ðhi þ 1 2hi þ hi1 Þ
qg i 2gh2i ðDxÞ2
# ð11:61Þ
ðhi þ 1 hi1 Þ2 ðhi þ 1 hi1 Þ
ðzbx Þi :
4ðDxÞ2 Dx
The computed water surface and bottom pressure profiles are compared in Fig. 11.7
with the corresponding test data (Sivakumaran 1981; Sivakumaran et al. 1983),
resulting in excellent agreement.
11.2 Two-Dimensional Steady Potential Flow 453
Fig. 11.7 Comparison of computed and measured (Sivakumaran 1981; Sivakumaran et al. 1983)
free surface and piezometric bottom pressure head in transcritical flow over a weir; a scaled photo
from experiments (Sivakumaran 1981) is also shown
454 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Paul Mansour Naghdi was born on March 29, 1924, in Tehran, Iran, and
passed away on July 9, 1994, aged 70 years at Berkeley CA, USA, having
been naturalized there in 1948. He made studies at the University of
Michigan, Ann Arbor MI, receiving the PhD degree in 1951. He was there
assistant until 1958, then associate and full professor. He moved to the
University of California, Berkeley CA, as a professor of engineering science,
chairing from 1964 to 1969 its Department of Applied Mechanics. He was a
member of the National Academy of Engineering NAE, the American
Society of Mechanical Engineers ASME, and ASME Fellow from 1969. He
was the 1958 Guggenheim Fellow, recipient of the Timoshenko Medal in
1980, and was awarded ASME Honorary Membership in 1983.
Naghdi’s work on continuum mechanics extended over forty years
including most aspects of the mechanical behavior of solids and fluids. He
was strongly attracted by fundamental questions and always sought to treat
these at the highest level of generality. He is known for his works in the areas
of shell theory and plasticity, but also in viscoelasticity, fluid sheets and jets,
continuum thermodynamics, and the mixture theory. His talents as a teacher
were legendary. At Berkeley, he continually refined a magnificent series of
courses on theoretical mechanics. These were well prepared, clearly deliv-
ered, original, and intellectually provocative. They also reflected his deep
understanding of the history of mechanics and his encyclopedic knowledge of
literature.
Transitional flows from mild to steep bottom slopes (Fig. 11.4) involve a continuous
free surface profile and a significant departure of the bottom piezometric pressure
profile from the free surface (Montes 1994; Castro-Orgaz and Hager 2009).
11.2 Two-Dimensional Steady Potential Flow 455
The numerical model developed in the previous section was applied to the experi-
mental data of Hasumi (1931) for a slope transition composed of a horizontal reach
followed by a circular-shaped transition profile of R = 0.1 m that finishes in a steep
slope reach of 45° inclination. The unit discharge is 0.0995 m2/s. The up- and
downstream boundary sections were located away from the slope break using
Dx = 0.005 m and 180 nodes to model the flow. The upstream boundary flow depth
was taken as the critical depth for hydrostatic flow hc = (q2/g)1/3, resulting in the total
head of the potential flow problem of H = 3hc/2. At the downstream boundary
section, the supercritical non-hydrostatic flow depth for streamlines nearly parallel to
the bed was estimated to hd = [(1 + S2o) q2/(2gE)]1/2, where E is the specific energy
at the downstream section and So = 1 is the chute slope for this test (Castro-Orgaz and
Hager 2009). The success in the iteration process depends upon the selection of a
plausible initial flow profile. In this test case, the initial flow profile was taken as
a rough linear interpolation between the critical depth and the brink flow depth
hb = 0.7hc along the horizontal reach, and between hb and hd along the chute portion.
The code for this test is available on the file Matthew_slopebreak.xls, in Chap. 12.
The bed profile derivatives were obtained using second-order central
finite-differences. The computed water surface and bed pressure profiles are com-
pared in Fig. 11.8 with the experimental data, resulting in a good agreement. Note the
computed pressure peak near the end of the transition circle, which is in disagreement
with the experiments, resulting from the downstream bed curvature discontinuity.
This computational effect can be removed by substituting the real bed profile by an
approximate transition curve with a smooth curvature variation (Castro-Orgaz and
Hager 2009). However, this task is not accomplished here, given that the emphasis is
on the computational process rather than on improving the simulation results.
Zerihun (2004) conducted steady non-hydrostatic flow tests for transcritical flow
over trapezoidal profiled weirs. The steady numerical model was applied to a
trapezoidal weir of up- and downstream slopes 2:1 (H:V) and crest width and height
both of 0.15 m. The unit discharge is 0.06128 m2/s. The up- and downstream
boundary sections were located away from the weir using Dx = 0.01 m and 300
nodes. The upstream boundary flow depth h1 was determined by iteration until the
upstream undulations in the approach flow were significantly suppressed. At the
downstream boundary section, the supercritical alternate depth of the upstream
depth was settled in the mathematical model. For the initial free surface profile,
linear variations along the weir involving the inflow depth, critical depth, brink
depth, supercritical depth at the toe of the weir, and the tailwater flow depth were
assumed. The code for this test is available on the file Matthew_embankmentweir.
xls, in Chap. 12. The bed profile derivatives were obtained using second-order
central finite-differences. The computed water surface and bed pressure profiles
after the iteration of h1 are compared in Fig. 11.9a with the experimental data
(Zerihun 2004), resulting in a good agreement. The pressure peaks at the slope
discontinuities can be removed using transition curves (Zerihun 2004), but this task
is not accomplished here.
A simulation for the same weir for q = 0.07102 m2/s using Dx = 0.005 m is
presented in Fig. 11.9b, showing again a good match with experimental observa-
tions. Lesleighter et al. (2008) simulated this weir test with CFD solving the
Reynolds-Averaged Navier–Stokes equations, and their results for the bottom
pressure head are presented in Fig. 11.9c. Note that the experimental data of
Zerihun (2004) did not reveal pressure peaks at the bed-slope breaks, but CDF
simulations highlight those peaks, although they are of smaller magnitude than
predicted by the 1D model. It is impossible to know the experimental magnitude of
these peaks, given that no pressure taps were installed at these positions (Zerihun
2004). However, the physical model experiments conducted by Lesleighter et al.
(2008) on trapezoidally profiled spillways reveal pressure peaks at slope breaks.
@h
þ DivðuhÞ ¼ 0: ð11:62Þ
@t
Zzs Zzs
1 1
uðx; y; z; tÞ U ðx; y; tÞ ¼ udz; vðx; y; z; tÞ V ðx; y; tÞ ¼ vdz: ð11:63Þ
h h
zb zb
@ @
wðx; y; z; tÞ ¼ ½U ðz zb Þ ½V ðz zb Þ ¼ Div½uðz zb Þ
@x @y
ð11:64Þ
@zb @zb
¼ wb gðDivuÞ; wb ¼ U þV ¼ u Gradðzb Þ;
@x @y
where η = z – zb and wb is the vertical velocity at the bed. Within this level of
mathematical approximation, boundary layers are neglected, and a slip velocity at
the bed is accepted for depth-averaged modeling purposes. If we replace u and
v everywhere by U and V in the general depth-integrated (x, y) momentum equa-
tions of Castro-Orgaz and Hager (2017), and turbulence stresses are neglected, then
the resulting inviscid equations can be expressed as a system of partial differential
equations in general conservative form as (Castro-Orgaz and Hager 2017) [see
Eqs. (1.24), (1.35) and (1.36)]
@U @F @G
þ þ @y0¼ S;
0 1 0
@t
1
@x 1
h Uh Vh 0 1
Rzs 0
B C B 2 B C
U ¼ @ Uh A; F ¼ @U hþ
1
pdz C
A; G¼B
VUh
Rzs C; S¼
@zb
q1 @ pb @x A:
q @ 2 A
zb V h þ q pdz
1
pb @z b
Vh UVh zb @y
ð11:65Þ
11.3 Unsteady Ideal Fluid Flow 459
In these equations, U is the dependent variable vector, F and G are the fluxes in the
x- and y-directions for non-hydrostatic pressure conditions, respectively, and S is
the source term. The pressure distribution p(z) is needed for model closure. The
vertical pressure distribution for depth-independent horizontal velocity components
and inviscid flow is given by (Castro-Orgaz et al. 2015) [see Eq. (1.60)]
2 z 3 2 z 3
Zzs Zs Zs
@ @ 4 @
pðx; y; z; tÞ ¼ qgðh gÞ þ q wdz þ q U wdz5 þ q 4V wdz5 qw2
@t @x @y
z z z
@I
¼ qgðh gÞ þ q þ qDivðIuÞ qw2 :
@t
ð11:66Þ
and
in which @h=@t has been replaced by −Div(uh) via Eq. (11.62). The vertical
pressure distribution can now be evaluated from Eq. (11.66) collecting
Eqs. (11.68)–(11.72). The result after some manipulation is
2
p 2 @u h g2
¼ gðh gÞ þ ½DivðuÞ Div u Grad½DivðuÞ
q @t 2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
f1
ð11:73Þ
@u
þ Gradðzb Þ þ u Grad½u Gradðzb Þ ðh gÞ:
@t
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
f2
Zh
p h2 h3 h2
dg ¼ g þ f1 þ f2 ; ð11:76Þ
q 2 3 2
0
Inserting Eqs. (11.76)–(11.77) into Eqs. (11.65), the general 2D inviscid Serre or
fully non-linear Boussinesq equations are obtained. These equations are used to model
fully non-linear and weakly dispersive water waves. These results presented here are in
agreement with those obtained by Nadiga et al. (1996) using a perturbation method.
For 1D water waves propagating over arbitrary topography, Eqs. (11.65) and
(11.73) read
@U @F
þ ¼ S;
@t @x
h Uh 0
U¼ ; F¼ 2 ; S ¼ @z b
;
Uh U h þ 2 gh þ 3 f1 h þ 2 f2 h
2 1 2 1 3 1 @x gh þ f1 h2 þ f2 h
ð11:78Þ
the undular bore (Peregrine 1966, 1967, 1972) (Fig. 11.11). The equations are
discussed by Barthelemy (2004), Cienfuegos et al. (2006), Dias and Milewski
(2010), and Bonneton et al. (2011) for coastal engineering problems.
Equations (11.78)–(11.79), the 1D inviscid Serre equations for weakly dispersive
and fully non-linear water waves over topography, were derived and numerically
solved by Seabra-Santos et al. (1987). The x-momentum equation in Eq. (11.78)
can be rewritten as
@ ðUhÞ @M pb @zb
þ ¼ ; ð11:80Þ
@t @x q @x
Note that M = gS. If unsteadiness is dropped, e.g., @ðÞ=@t and U = q/h is used to
transform the U-derivatives into h-derivatives, Eqs. (11.81)–(11.82) reduce to
11.3 Unsteady Ideal Fluid Flow 463
Eqs. (11.22) and (11.21), respectively. This means that the steady version of the
SGN equations derived here equals the steady Picard iteration model. The imme-
diate implication is that Eq. (11.19), an extended Bernoulli-type equation, is the
integral form of Eq. (11.78) in steady state. It has important applications for water
wave flow problems. It is common practice while developing unsteady flow solvers
of the SWE to test their ability to converge to steady flow solutions. The conver-
gence to steady transcritical flow over a weir is a common and widely accepted test
(see Sect. 9.9.5). This issue was exported to non-hydrostatic Boussinesq-type sol-
vers forcing a version of these models with the dispersive terms deactivated to fight
with this test and to demonstrate their convergence to the steady flow solution.
However, it is more logic and stringent to test the ability of a non-hydrostatic
unsteady flow solver to converge to the correct steady non-hydrostatic flow solu-
tion. This opens the path of using Eq. (11.19) as a means to generate “exact” steady
non-hydrostatic flow solutions and to test the ability of non-hydrostatic solvers of
Eq. (11.78) to converge to these solutions. We remark then that the relevance of the
steady flow solver of Eq. (11.19) is therefore not only its ability to reveal the role
on non-hydrostaticity in weir flow, but also their utility to generate steady flow
solutions of the more general unsteady flow system.
For numerical computations, it is desirable to collect in a single vector all the
terms with temporal derivatives. This task is accomplished here. Consider the
following identity
@ h3 @U h3 @ 2 U @h @U
¼ þ h2 ; ð11:83Þ
@t 3 @x 3 @t@x @t @x
Using this identity, it is possible to split terms with temporal derivatives. Following
the same strategy for the remaining terms with time derivatives in M and pb, the
following alternative form of Eq. (11.78) is obtained
@W @F
þ ¼ S þ Sd ;
@t @x
h Uh
W¼ ; F¼ ; ð11:85Þ
r U 2 h þ 12 gh2
@zb 0 @zb 0 @D 0
S¼ ; Sd ¼ ;
@x gh @x p1 @x 1
where
464 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
1 3 @2U 2 @U @h @zb @zb @h 1 2 @ 2 zb
r ¼ Uh h h þ Uh þ þ Uh ; ð11:86Þ
3 @x2 @x @x @x @x @x 2 @x2
" #
@U 2 @2U 1 3 @U @zb 2 @ zb 1 2
2
D¼ U 2 h þ U þU h
@x @x 3 @x @x @x2 2
ð11:87Þ
@ ðUhÞ @zb @U
þ U h h;
@x @x @x
" #
@U 2 @2U 2 @U @zb 2 @ zb
2
p1 ¼ U 2 h þ U þU h
@x @x @x @x @x2
ð11:88Þ
@ ðUhÞ @zb @U
þ U h :
@x @x @x
Consider waves of permanent form propagating over still water with celerity
c. Using the Galilei transformation X = x − ct for a wave displacement in the
positive x-direction over a horizontal bed, the Serre–Green–Naghdi equations,
Eq. (11.78), take the form (Castro-Orgaz and Hager 2017)
@h @ðUhÞ
c þ ¼ 0; ð11:89Þ
@X @X
@ðUhÞ @ 1 2 1
c þ gh þ U 2 h þ UX2 UUXX þ cUXX h3 ¼ 0: ð11:90Þ
@X @X 2 3
The solitary wave is a solution of Eq. (11.92) subject to the boundary conditions
h ! ho and U ! 0 for x ! ±∞, with ho as still water depth. Integrating twice, the
1=2
final result is with Fo ¼ q= gh3o (Serre 1953; Benjamin and Lighthill 1954;
Castro-Orgaz and Hager 2017)
" 1=2 #
hðx; tÞ 3F2o 3 ðx ctÞ
¼ 1 þ F2o 1 sech2 : ð11:93Þ
ho Fo 2ho
The maximum flow depth at the solitary wave crest is obtained from Eq. (11.93) at
x − ct = 0 as
c2 h2o c2
hmax ¼ ho F2o ¼ ho ¼ : ð11:96Þ
gh3o g
For given values of ho and A, the degree of non-linearity of the solitary wave is
fixed. The celerity c and Fo are thus deduced from Eqs. (11.97)–(11.98), and,
resorting to Eqs. (11.93) and (11.95), the functions h(x, t) and U(x, t) are fully
466 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Joseph Valentin Boussinesq was born on March 15, 1842, at St. André-
de-Sangonis, France, and passed away on February 19, 1929, in Paris. He was
self-taught, starting his scientific writing in 1865. He thereby took into
consideration during his long career all branches of mathematical physics
except for electro-magnetism. After having served as teacher at various
colleges of France, he was appointed in 1873 Lecturer at the University of
Lille. In 1886, Boussinesq was appointed to the chair of mechanics at the
famous Sorbonne University, Paris, taking over in 1896 as professor of
mathematical physics at Collège de France.
11.3 Unsteady Ideal Fluid Flow 467
The system of Eq. (11.85) is solved here using a finite volume-finite difference
method based on the MUSCL-Hancock scheme, a second-order accurate model in
space and time (see Sect. 9.7). Boussinesq-type water wave propagation models are
solved for coastal engineering applications using fourth-order accurate schemes in
space, and third-order accuracy or fourth-order accuracy in time. The reason for
imposing such accuracy is that truncation errors originating from the discretization
to second-order accuracy of the Saint-Venant type leading terms can induce
numerical dispersion. This problem is serious for large-scale simulations using
sparse time-space meshes (Wei et al. 1995). However, using a fine mesh in
second-order accurate schemes, this effect disappears or is significantly limited.
There are good reasons for testing a second-order non-hydrostatic scheme, how-
ever: The current knowledge of the finite volume technology for the solution of the
SWE is in a state of maturity permitting the development of a large variety of 1D
and 2D solvers worldwide. Research groups developed their own codes and verified
them extensively. If the inclusion of non-hydrostatic effects into a solver is offered
468 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Dt
Wadv ¼ Wki Fi þ 1=2 Fi1=2 þ DtSi : ð11:99Þ
i
Dx
Here, Δt and Δx are the step sizes in the x and t axes, respectively, k refers to the
time level, i is the cell index in the x-direction, and Fi+1/2 is the numerical flux
crossing the interface i + 1/2 between cells i and i + 1. Note that the advection step
as defined here includes the bed-slope source term directly. The topography source
term is discretized for a well-balanced scheme as [see Eq. (9.99)]
@zb 0 hi þ 1=2 þ hi1=2 zbi þ 1=2 zbi1=2 0
Si ¼ g h ¼ g: : ð11:100Þ
@x i 1 2 Dx 1
The solution process starts with the cell-averaged values of conserved variables at
time level k, Uki . For second-order space accuracy, a piecewise linear reconstruction
is conducted within each cell (Toro 2001). Linear slopes resulting from the
reconstructed solution must be limited to avoid spurious oscillations near discon-
tinuities. Let letters L and R denote the reconstructed variables at the left and right
sides of a cell interface; the resulting values of U at each of its sides are with DUi
and DUi þ 1 as the limited jumps (Toro 2001) [see Eqs. (9.122)–(9.125)]
1 1
ULi þ 1=2 ¼ Uki þ DUi ; URi þ 1=2 ¼ Ukiþ 1 Uki DUi þ 1 : ð11:101Þ
2 2
The surface gradient method (SGM) (Zhou et al. 2001) is adopted, consisting of
the reconstruction of the free surface elevation, instead of the water depth h. The
minmod limiter is used in the tests presented here. In the MUSCL-Hancock method,
an evolution of boundary-extrapolated values ULi þ 1=2 and URi þ 1=2 at interface i + 1/2
over half the time step is conducted to regain second-order accuracy in time. Based
on a Taylor series expansion in space and time, interface values are then given by
(Toro 2001) [see Eq. (9.134)]
11.3 Unsteady Ideal Fluid Flow 469
L Dt h L i Dt
Ui þ 1=2 ¼ ULi þ 1=2 F Ui þ 1=2 F URi1=2 þ Si ;
2Dx 2 ð11:102Þ
R Dt h L i Dt
Ui þ 1=2 ¼ Ui þ 1=2
R
F Ui þ 3=2 F URi þ 1=2 þ Si þ 1 :
2Dx 2
L R
With these evolved boundary-extrapolated variables Ui þ 1=2 and Ui þ 1=2 defining states
L and R, the numerical flux is computed using the HLL approximate Riemann
solver as (Toro 2001) [see Eq. (9.44)]
8
> FL if SL 0
>
<
SR FL SL FR þ SR SL ðUR UL Þ
Fi þ 1=2 ¼ ; if SL 0 SR : ð11:103Þ
>
> SR SL
:
FR if SR 0
SL ¼ UL cL kL ; SR ¼ UR þ cR kR ; ð11:104Þ
The flow depth at the star region of the Riemann problem at each interface h* is
(Toro 2001) [see Eq. (9.47)]
2
1 1 1
h ¼ ð cL þ cR Þ þ ð U L U R Þ : ð11:106Þ
g 2 4
For the dry-bed problem, the celerity of the signals is given by [see Eqs. (9.48)–
(9.49)]
2 3
6 Dx
Dt ¼ CFL4
7
5: ð11:108Þ
k 1=2
max
Ui þ ðghki Þ
Once the solution of Eq. (11.99) is available, the value obtained for the flow depth
is hk+1; the auxiliary variable r, however, must be updated to include the effect of
Sd. Here, we will use a predictor–corrector finite-difference scheme to incorporate
Sd in the solution. First, a predictor step is conducted as
@zb @D adv
rpi ¼ radv þ Dt p 1 : ð11:109Þ
i
@x @x i
All the spatial derivatives in the non-hydrostatic source term (linked to p1, D, and
its gradient) are approximated using second-order central finite-differences. In
general, a derivative is discretized in this work as
@mf 1 Xþk
¼ x k fk ; ð11:110Þ
@xm i Dxm k
where the weights are given in Table 11.2. The only exception is the first derivative
of U, which is discretized in the predictor step using a forward finite-difference as
(Mohapatra and Chaudhry 2004)
Ui þ 1 Ui
Ux ¼ : ð11:111Þ
Dx
Once rpi is determined, the predicted velocity field must be resolved. This is
accomplished by solving an elliptic problem posed by Eq. (11.86). The process is
explained below, given that it must be also applied after the correction step.
The corrector step is given by
@zb @D p
rki þ 1 ¼ radv þ Dt p 1 ; ð11:112Þ
i
@x @x i
Table 11.2 Weighting factors for discretization of spatial derivatives in non-hydrostatic terms
based on Abramowitz and Stegun (1972) and Fornberg (1988)
Type of Order of Order of Weighting factor xk at nodes
derivative derivative accuracy k = −2 k = −1 k = 0 k =+1 k = +2
Centered 1 2 0 −1/2 0 +1/2 0
Centered 1 4 1/12 −2/3 0 2/3 −1/12
Centered 2 2 0 1 –2 1 0
Centered 2 4 −1/12 4/3 –5/2 4/3 −1/12
Upwind 1 1 – −1 1 – –
Upwind 1 2 1/2 −2 3/2 – –
11.3 Unsteady Ideal Fluid Flow 471
which is adopted as the final step. The first derivative of U is discretized in the
corrector step with a backward finite-difference as (Mohapatra and Chaudhry 2004)
Ui Ui1
Ux ¼ : ð11:113Þ
Dx
Once the values of r are determined at each finite volume for the new time level,
the following elliptic problem is stated at each cell of the computational domain
rki þ 1 ¼ d1 Ui1
kþ1
þ d2 Uik þ 1 þ d3 Uikþþ11 ; ð11:114Þ
h3i h2i
d1 ¼ 2
þ ðhi þ 1 hi1 Þ;
3ðDxÞ 4ðDxÞ2
2h3i
d2 ¼ hi þ þ a;
3ðDxÞ2
h3i h2i
d3 ¼ 2
ðhi þ 1 hi1 Þ;
3ðDxÞ 4ðDxÞ2
ðzbi þ 1 zbi1 Þ ðzbi þ 1 zbi1 Þ ðhi þ 1 hi1 Þ ðzbi þ 1 2zbi þ zbi1 Þ
a ¼ hi þ þ h2i :
2Dx 2Dx 2Dx ðDxÞ2
ð11:115Þ
qk1 þ 1 ¼ qinlet ;
ð11:116Þ
hk1 þ 1 ¼ hk2 þ 1 :
qkNþ 1 ¼ qkN1
þ1
;
ð11:117Þ
hkNþ 1 ¼ hkN1
þ1
:
13. Once rk+1 is available at all cells, the tridiagonal system is solved using the
Thomas algorithm to obtain the final velocities [Eq. (11.114)].
14. If the actual time is equal to the final time, then stop.
15. Go back to step 3 for a new time loop.
In this section, a high-resolution scheme in both space and time O(Dt3, Dx4) is
developed to assess the second-order MUSCL-Hancock scheme. For the time
stepping, a third-order strong stability preserving (SSP) Runge–Kutta scheme is
given by (Gottlieb et al. 2001)
h i
ð1Þ ð1Þ ð1Þ ð1Þ
Wi ¼ Wki þ L Uki Dt þ ðSd Þi Dt ) Ui ¼ E 1 Wi ;
h i h i
ð2Þ ð1Þ ð1Þ ð2Þ ð2Þ ð2Þ
Wi ¼ Wi þ L Ui Dt þ ðSd Þi Dt ) Ui ¼ E 1 Wi ;
h i
ð3Þ ð2Þ ð3Þ ð3Þ
Wi ¼ 34 Wki þ 14 Wi ) Ui ¼ E 1 Wi ; ð11:118Þ
h i h i
ð4Þ ð3Þ ð3Þ ð4Þ ð4Þ ð4Þ
Wi ¼ Wi þ L Ui Dt þ ðSd Þi Dt ) Ui ¼ E 1 Wi ;
h i
ð4Þ ðk þ 1Þ
Wki þ 1 ¼ 13 Wki þ 23 Wi ) Uki þ 1 ¼ E1 Wi :
Here, E[ ] is the elliptic operator linked to Eq. (11.114), and L[] is the finite
volume-finite difference spatial operator
1
L½Ui ¼ Fi þ 1=2 Fi1=2 þ Si : ð11:119Þ
Dx
In the space operator L(), U is reconstructed with fourth-order accuracy, but the
elliptic operator E() is maintained with second-order accuracy to preserve the
tridiagonal structure of the linear system of equations determining the
non-hydrostatic velocity field. Sd must be updated with U at the corresponding time
stage, so iteration is needed. A high-order total variation diminishing monotone
upstream centred scheme for conservation laws (MUSCL-TVD-4th) is adopted to
reconstruct the solution (Erduran et al. 2005). The local Riemann problem at each
cell interface is then determined by the vector U at its left (L) and right (R) sides
from
1 1
ULiþ 1=2 ¼ Ui þ uðr1 ÞD Ui1=2 þ 2u D Ui þ 1=2 ; ð11:120Þ
6 r1
474 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
1 1
URiþ 1=2 ¼ Ui þ 1 2uðr2 ÞD Ui þ 1=2 þ u D Ui þ 3=2 ; ð11:121Þ
6 r1
1
D Ui þ 1=2 ¼ DUi þ 1=2 DUi þ 3=2 2DUi þ 1=2 þ DUi1=2 ; ð11:122Þ
6
ri þ j ri j D Ui þ 1=2 D Ui þ 3=2
uðri Þ ¼ ; r1 ¼ ; r2 ¼ : ð11:128Þ
1 þ j ri j D Ui1=2 D Ui þ 1=2
The surface gradient method is applied to reconstruct the water surface elevation.
Once the reconstruction step is finished, the numerical flux Fi+1/2 is estimated with the
HLL approximate Riemann solver. The dispersive source term Sd (with the exception
of Ux) is discretized using fourth-order accurate central finite-differences (Table 11.2)
for the cells i = 3 to N – 2. At the cells i = 2 and N – 1, Sd is discretized using
second-order differences, and dispersive effects are deactivated at the ghost cells i = 1
and i = N (Sd = 0) to avoid any non-hydrostatic influence transmitted to the cells
i = 2 and i = N − 1 through the gradient @D=@x = (Di+1−Di−1)/(2Δx). The
derivative Ux is computed using upwind differencing of first order at cells i = 2 and
N – 1, and of second order at cells i = 3 to N – 2 (Table 11.2).
The process of numerical solution is as follows:
1. A longitudinal finite volume mesh with cell width Dx is defined and the bed
profile function zb(x) is used to define the bed elevation at the cell faces. The
bed elevation at the cell center is taken as the average of cell faces, consistent
with a linear variation of the bed profile within the cell. The bed derivatives for
the dispersive terms are evaluated using the computed coordinates at cell
centers by second-order central finite-differences at cells i = 2 and N – 1, and
with fourth-order central differences at the other cells.
11.3 Unsteady Ideal Fluid Flow 475
Figure 11.12 displays the unsteady wave evolution for the symmetrical hump of
shape zb = 0.2exp[−0.5(x/0.24)2] (m) according to Sivakumaran et al. (1983). The
unit discharge is 0.11197 m2/s. The up- and downstream boundary sections are
located at x = ± 3 m. The mesh used is of Δx = 0.01 m with CFL = 0.4. For
comparative purposes, the same simulation was conducted solving the SWE. For
the simulations shown in Fig. 11.12, a dam break numerical setup is implemented;
a gate is located at coordinate x = 1 m. The water is static, with free surface
elevations of 0.34 and 0.03 m up- and downstream of the gate, respectively. The
gate is removed instantaneously at time t = 0; the steady discharge is introduced at
the inlet section as a pulse and kept constant during the simulation. The numerical
model used is implemented in a code available on the file “MUSCLHancock_weir.
xls,” in Chap. 12.
At time t = 0, the dam break initial conditions generate a downstream traveling
dispersive shock wave and an upstream traveling rarefaction wave (Fig. 11.12a, b).
Compare the SGN and the SWE, where a solitary wave at the shock front is
predicted by the former. At the upstream section, the addition of the discharge pulse
generates another wave. In Fig. 11.12c, an undular wave pattern upstream of the
weir is visible, and a dam break front with undulations according to the SGN
equations. The SWE predict a sharp shock front, as expected from the second-order
MUSCL-Hancock scheme. In Fig. 11.12d, the upstream wave train is still active,
476 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Fig. 11.12 Evolution of water surface profile h(x) at various times t in unsteady weir flow
[adapted from Castro-Orgaz and Cantero-Chinchilla (2019)]
11.4 Unsteady Flow Test Cases 477
while supercritical flow develops along the downstream weir face, with a hydraulic
jump near the toe of the weir. The undular waves of the dam break front are
progressively leaving the computational domain. In Fig. 11.12e the upstream water
waves continue their activity looking for equilibrium in the upstream water depths,
while the hydraulic jump phenomena increase their intensity at the downstream side
of the obstacle. Dispersive effects gain weight, and a large amplitude initial wave is
formed at the jump front, as is typical for undular hydraulic jumps. Figure 11.12f
shows that the upstream waves start to diminish in wave amplitude, and that the
undular hydraulic jump is pushed out of the computational domain by the super-
critical flow on the downstream weir face. Figure 11.12g finally shows the
steady-state results obtained from the simulation.
Before discussing the steady flow results, the adequacy of the implemented
transmissive boundary condition is assessed in Fig. 11.13. Here, the computational
478 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
results are shown for t = 2.5 s using a domain of double length downstream of the
weir crest. Note that the waves are freely leaving the computational domain without
significant alterations. A more accurate approach would require the inclusion of
sponge layers at both ends of the domain (Cantero-Chinchilla et al. 2018).
The steady flow results generated by the unsteady flow solver are compared in
Fig. 11.14 with the “exact” steady flow solution previously presented in Fig. 11.7.
Note that the two simulations are in excellent agreement both for the free surface
and the bed pressure profiles. The bed pressure head for the unsteady flow solver
was determined using a finite-difference discretization of Eq. (11.82) based on the
computed vector U at each time t. Therefore, the proposed generalization of the
MUSCL-Hancock scheme of the SWE is able to deal with this problem introducing
the simple implementation of an additional source term and a tridiagonal equation
solver.
It is common practice in water wave simulations for ocean research to use
high-resolution schemes, typically of order O(Dt3, Dx4). This is done to reduce
truncation errors and to avoid numerical dispersion. A question may be well asked
if this is also necessary while producing steady non-hydrostatic flow solutions. It
is shown in Fig. 11.14 that the O(Dt2, Dx2) scheme produces a highly accurate
steady flow solution. Now, we test if this accuracy is maintained using a coarse
mesh. A new unsteady flow simulation was conducted dividing the numbers of
computational cells in Fig. 11.15, e.g., using a mesh of only Δx = 0.1 m, and
imposing the maximum time step for stability by trial and error, which was found at
CFL = 0.7. The results shown in Fig. 11.15a reveal that even for this coarse mesh,
results are still very good. The conclusion is that second-order accuracy is enough
for the simulation of steady non-hydrostatic flows, and centred discretizations
produce highly accurate results both for steady and unsteady solvers. Finally,
Fig. 11.15b presents the results for Δx = 0.2 m and CFL = 0.25. With this mesh,
the total number of cells is 30, and the flow over the obstacle is represented only by
9 of these cells. There is an obvious loss of accuracy for this extremely coarse mesh,
but the prediction is still in fair agreement with the exact results.
A comparison of the MUSCL-Hancock scheme with the third-order RK/
fourth-order MUSCL scheme at t = 1 s and 2.5 s is displayed in Fig. 11.16, using
the same mesh adopted to produce Fig. 11.12. The two are in good agreement, with
some deviations at the undular jump on the tailwater, indicating that the mesh
adopted for MUSCL-Hancock simulation is adequate, but should be refined to get
more precise results. The high-resolution model is implemented in a code available
on the file “rk3MUSCL4_weir.xls,” in Chap. 12. In general, a possible practical
480 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Fig. 11.16 Comparison of MUSCL-Hancock method with the rk3MUSCL4 solver during
transient flow
strategy for using an O(Dt2, Dx2) SWE scheme improved with dispersive terms is to
run a first simulation, and, if extreme waves are expected in any portion of the
computational domain, to refine locally the mesh there to increase the model res-
olution (Popinet 2015).
The solitary wave [Eq. (11.93)] is a solution of the SGN equations in which
non-linear and dispersive effects are balanced. Non-linear effects are present in the
SWE, but this model is dispersionless. The SGN are fully non-linear and weakly
dispersive Boussinesq equations. Thanks to the full non-linearity in the
non-hydrostatic terms, there is no limitation on the wave amplitude modeled by
SGN equations. However, the equations are weakly dispersive, and, thus, only long
waves are correctly propagated. A solitary wave of large amplitude is therefore a
possible theoretical solution within the domain of validity of the SGN equations.
However, exact balancing of non-linear and dispersive effects is necessary to
produce a wave of permanent form traveling over still water. As the solitary wave is
an analytical solution of Eq. (11.78) resulting from such balancing, a numerical
11.4 Unsteady Flow Test Cases 481
Fig. 11.17 Propagation of solitary wave of A/ho = 0.2 and ho = 1 m using the MUSCL-Hancock
scheme at t = 20 s
scheme to solve them must be able to preserve this equilibrium while propagating
an analytical solitary wave as input data. The ability of the MUSCL-Hancock
scheme to propagate a solitary wave is investigated in Fig. 11.17. A test case
involving ho = 1 m and A = 0.2 m (Fig. 11.17a) is considered in Fig. 11.17b–e at
t = 20 s after initiation of routing. The numerical model is implemented in a code
available on the file “MUSCLHancock_solitarywave.xls,” in Chap. 12.
A comparison between the analytical and numerical solutions as functions of
Δx and CFL indicates that if progressively reducing both, it is possible to find a
mesh where the numerical prediction perfectly matches the theoretical expression
482 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Fig. 11.18 Comparison of a rk3MUSCL4 solver with b MUSCL-Hancock solver for solitary
wave propagation
(Fig. 11.17e). The MUSCL-Hancock scheme for the SGN equations is thus able to
preserve the balance between non-linear and dispersive effects. This is not the case
for the same scheme applied to the SWE (Fig. 11.17f), given that dispersive effects
are absent. After 20 s of routing, the SWE transform the input solitary wave into a
shock.
A test case involving ho = 1 m and A = 0.5 m is considered in Fig. 11.18 at
t = 15 s after initiation of routing. A mesh with Δx = 0.15 m and CFL ¼ 0:8 was
used in the rk3MUSCL4 numerical model, with the results shown in Fig. 11.18a.
The numerical model is implemented in a code available on the file
“rk3MUSCL4_solitarywave.xls,” in Chap. 12. Note that the high-resolution model
produces an excellent result. The same test with the same mesh was implemented in
the MUSCL-Hancock scheme, with the results shown in Fig. 11.18b. There is an
obvious distortion of the solitary wave profile for this mesh, indicating the need to
refine the mesh in the O(Dt2, Dx2) scheme. This comparison was conducted for
a high-amplitude wave, A/ho = 0.5, which clearly reveals the benefits of an
O(Dt3, Dx4) scheme for this extreme non-linearity.
The balance between non-linearity and dispersion in non-hydrostatic waves can
experimentally be observed in surges. An abrupt gate closure generates a surge
propagating in the upstream direction (Rouse 1961). If the wave amplitude is much
less than the final depth, the surge is undular (Fig. 11.19a, b) (Rouse 1961), from
where the quoted balance between non-linear and dispersion effects becomes
apparent. If the amplitude increases, there will be a state where the first wave
breaks, but there is still an undular wave in the tailwater (Fig. 11.19c, d). If the
wave amplitude is much higher than the final depth, a distinctly turbulent broken
wave is generated (Fig. 11.19e, f) (Rouse 1961). The SWE are able to simulate
broken surges, whereas the SGN equations can simulate undular waves. A criterion
for wave breaking is needed to switch from the SWE to the SGN equations and thus
predict the complete wave evolution.
11.4 Unsteady Flow Test Cases 483
Fig. 11.24 Non-hydrostatic dam break wave for hd =ho [ 0:5 (photos by O. Castro-Orgaz at
hydraulic flume of University of Córdoba) a Initial condition b Undular wave front c Reflected
undular wave front
References
Abramowitz, M., & Stegun, I. A. (1972). Handbook of mathematical functions with formulas,
graphs, and mathematical tables (10th ed.). New York: Wiley.
Barthelemy, E. (2004). Nonlinear shallow water theories for coastal waters. Surveys In
Geophysics, 25(3), 315–337.
Benjamin, T. B., & Lighthill, M. J. (1954). On cnoidal waves and bores. Proceedings of the Royal
Society London, A, 224, 448–460.
Blau, E. (1963). Der Abfluss und die hydraulische Energieverteilung über einer parabelförmigen
Wehrschwelle [Distributions of discharge and energy over a parabolic-shaped weir].
Mitteilungen der Forschungsanstalt für Schiffahrt, Wasser- und Grundbau, Berlin, Heft 7,
5–72 (in German).
Bonneton, P., Barthelemy, E., Chazel, F., Cienfuegos, R., Lannes, D., Marche, F., et al. (2011).
Recent advances in Serre-Green-Naghdi modelling for wave transformation, breaking and
runup processes. European Journal of Mechanics B/Fluids, 30(6), 589–597.
Boussinesq, J. (1877). Essai sur la théorie des eaux courantes [Memoir on the theory of flowing
water]. Mémoires présentés par divers savants à l’Académie des Sciences, Paris 23, 1–660; 24,
1-60 (in French).
490 11 Numerical Modeling of Non-hydrostatic Free Surface Flows
Cantero-Chinchilla, F. N., Castro-Orgaz, O., Dey, S., & Ayuso, J. L. (2016). Nonhydrostatic dam
break flows I: Physical equations and numerical schemes. Journal of Hydraulic Engineering,
142(12), 04016068.
Cantero-Chinchilla, F. N., Castro-Orgaz, O., & Khan, A. A. (2018). Depth-integrated nonhydro-
static free-surface flow modelling using weighted-averaged equations. International Journal
for Numerical Methods in Fluids, 87(1), 27–50.
Castro-Orgaz, O., & Cantero-Chinchilla, F. N. (2019). Non-linear shallow water flow over
topography with depth-averaged potential equation. Environmental Fluid Mechanics, in Press.
Castro-Orgaz, O., & Chanson, H. (2016). Closure to Minimum specific energy and transcritical
flow in unsteady open channel flow. Journal of Irrigation and Drainage Engineering, 142(10),
07016015.
Castro-Orgaz, O., & Hager, W. H. (2009). Curved streamline transitional flow from mild to steep
slopes. Journal of Hydraulic Research, 47(5), 574–584.
Castro-Orgaz, O., & Hager, W. H. (2013). Velocity profile approximations for two-dimensional
potential channel flow. Journal of Hydraulic Research, 51(6), 645–655.
Castro-Orgaz, O., & Hager, W. H. (2014). 1D modelling of curvilinear free surface flow:
Generalized Matthew theory. Journal of Hydraulic Research, 52(1), 14–23.
Castro-Orgaz, O., & Hager, W. H. (2017). Non-hydrostatic free surface flows. Advances in
Geophysical and Environmental Mechanics and Mathematics. 696 p. Berlin: Springer. https://
doi.org/10.1007/978-3-319-47971-2
Castro-Orgaz, O., Hutter, K., Giraldez, J. V., & Hager, W. H. (2015). Non-hydrostatic granular
flow over 3D terrain: New Boussinesq-type gravity waves? Journal of Geophysical Research:
Earth Surface, 120(1), 1–28.
Chanson, H. (2004). The hydraulics of open channel flows: An introduction. Oxford, UK:
Butterworth-Heinemann.
Cienfuegos, R., Barthélemy, E., & Bonneton, P. (2006). A fourth-order compact finite volume
scheme for fully nonlinear and weakly dispersive Boussinesq-type equations. Part I: Model
development and analysis. International Journal for Numerical Methods in Fluids, 51(11),
1217–1253.
Dias, F., & Milewski, P. (2010). On the fully non-linear shallow-water generalized Serre
equations. Physics Letters A, 374(8), 1049–1053.
Erduran, K. S., Ilic, S., & Kutija, V. (2005). Hybrid finite-volume finite-difference scheme for the
solution of Boussinesq equations. International Journal for Numerical Methods in Fluids, 49
(11), 1213–1232.
Fenton, J. D. (1996). Channel flow over curved boundaries and a new hydraulic theory.
Proceeding of 10th IAHR APD Congress, Langkawi (pp. 266–273). Malaysia.
Fornberg, B. (1988). Generation of finite difference formulas on arbitrarily spaced grids.
Mathematics of Computation, 51(184), 699–706.
Gottlieb, S., Shu, C.-W., & Tadmor, E. (2001). Strong stability-preserving high-order time
discretization methods. SIAM Review, 43(1), 89–112.
Green, A. E., & Naghdi, P. M. (1976a). Directed fluid sheets. Proceedings of the Royal Society.
London, A, 347, 447–473.
Green, A. E., & Naghdi, P. M. (1976b). A derivation of equations for wave propagation in water of
variable depth. Journal of Fluid Mechanics, 78, 237–246.
Hager, W. H. (1985). Critical flow condition in open channel hydraulics. Acta Mechanica, 54(3–4),
157–179.
Hager, W. H., & Hutter, K. (1984a). Approximate treatment of plane channel flow. Acta
Mechanica, 51(3–4), 31–48.
Hager, W. H., & Hutter, K. (1984b). On pseudo-uniform flow in open channel hydraulics. Acta
Mechanica, 53(3–4), 183–200.
Hasumi, M. (1931). Untersuchungen über die Verteilung der hydrostatischen Drücke an
Wehrkronen und -Rücken von Überfallwehren infolge des abstürzenden Wassers [Studies on
the distribution of hydrostatic pressure distributions at overflows due to water flow]. Journal
Department of Agriculture, Kyushu Imperial University 3(4), 1–97 (in German).
References 491
The Visual Basic language is used to write scripts, and these scripts are run as
“Macros” in Excel. Prior knowledge of VBA is necessary, however. Given that the
language is fairly simple, this is not a real limitation. Excel itself is used as a means to
organize the input data and receive the output from the code; graphics are immedi-
ately displayed with the computational results. Codes are written as clear as possible
showing all steps. The use of subroutines is done to structure the computations into
separate operations. Programs are by no means optimized; they are simply made to
avoid losing any intermediate step. Further, most codes are tailored for a specific
application, and, thus, are not to be used for general purposes. However, the reader
may take pieces of the codes to develop his/her own general-purpose code. Despite
the codes are written in VBA, their translation to FORTRAN is fairly simple, and the
material can be used by readers as the base to develop codes for scientific purposes.
The codes available as supporting material are listed in Table 12.1. Below, examples
of the codes for steady and unsteady flows are given. The unsteady flow VBA codes
are adapted from the FORTRAN code HW_MUSH.F by Toro (2000, 2001).
Q = Worksheets(''hoja1'').Cells(3, 4).Value
b = Worksheets(''hoja1'').Cells(4, 4).Value
z = Worksheets(''hoja1'').Cells(5, 4).Value
numero = Worksheets(''hoja1'').Cells(6, 4).Value
sections = Worksheets(''hoja1'').Cells(7, 4).Value
So = Worksheets(''hoja1'').Cells(8, 4).Value
n = Worksheets(''hoja1'').Cells(9, 4).Value
dx = Worksheets(''hoja1'').Cells(10, 4).Value
ho = Worksheets(''hoja1'').Cells(11, 4).Value
xo = Worksheets(''hoja1'').Cells(12, 4).Value
direc = Worksheets(''hoja1'').Cells(13, 4).Value
' boundary point
Worksheets(''hoja1'').Cells(7, 7).Value = xo
500 12 Numerical Library of Shallow Water Equations
Worksheets(''hoja1'').Cells(7, 8).Value = ho
' select direction of computation
If direc = 0 Then dx = −dx
h2 = ho
x = xo
' loop to go across all the channel sections
For j = 1 To sections
Newton
x = x − dx
h = h1
h2 = h1
Worksheets(''hoja1'').Cells(7 + j, 7).Value = x
Worksheets(''hoja1'').Cells(7 + j, 8).Value = h
Next j
End Sub
Sub Newton()
'Program to compute the unknown depth using Newton–Rapshon method
' Initiate variables
hsal = h2
section
Fo = So * dx – E − 0.5 * dx * Sf
h1 = h2
' Iterative computation of depth
For i = 1 To numero
hsal = h1
section
functions
hnew = h1 − f/dfdh
If Abs((hnew − h1) /h1) < 10 ^ −8 Then GoTo 10
h1 = hnew
Next i
12.2 Examples of Codes 501
10
End Sub
Sub section()
' variables of trapezoidal cross section
A = b * hsal + z * hsal ^ 2
BSUP = b + 2 * z * hsal
Froude = (Q ^ 2 /9.81 /A ^ 3 * BSUP) ^ 0.5
p = b + 2 * hsal * (1 + z ^ 2) ^ 0.5
R = A /p
E = hsal + Q ^ 2 /2 /9.81 /A ^ 2
Sf = n ^ 2 * Q ^ 2 /A ^ 2 /R ^ (4 /3)
End Sub
Sub functions()
' functions to apply the NR iteration method
f = E − 0.5 * dx * Sf + Fo
dEdh = 1 − Froude ^ 2
dSfdh = Sf * (−10 /3 * BSUP /A + 4 /3 /p * 2 * (1 + z ^ 2) ^ 0.5)
dfdh = dEdh − 0.5 * dx * dSfdh
End Sub
Public TOLDEP
Public i As Integer, k As Integer, count As Integer
Public N As Integer, Tmax, xdam
Sub Main()
' Program for the computation of Dam Break waves using the Finite Volume method
with: MUSCL linear reconstruction,
' minmod/superbee slope limiter, HLL approximate Riemann solver, and Euler time
stepping
' Second order accuracy is achieved with Hancock's method
' Simplified test conditions are:
' * Horizontal channel
' * Rectangular cross section
' * Dry and wet fronts allowed
' * Friction computed using Manning's equation
' * Higher water level is on the left, so flow is from left to right
' * Waves can not reach the boundary sections of the domain
' by Oscar Castro-Orgaz,
' This code is based on the FORTRAN code HW_MUSH.F by Prof. E. Toro (Numerica
library, 2000)
-15
TOLDEP = 10
lambda = Worksheets("Input").Cells(11, 4).Value
fric = Worksheets("Input").Cells(13, 4).Value
' Solve Riemann problem using the boundary extrapolated values of conserved
variables, after application
' of a limiter to avoid spurious oscillations
Riemann
' Compute cell-averaged conserved variables at new time level
evolution
' Check dry bed conditions
dry
' Include friction effects from source terms
' Implicit computation
If fric = 1 Then
friction
End If
' Explicit computation
If fric = 2 Then
friction2
End If
t = t + dt
count = count + 1
' Check final of transient computations
If t < Tmax Then GoTo 10
' Print water depth and discharge data
For i = 1 To N
Worksheets(''Input'').Cells(4 + i, 10).Value = x(i) − 0.5 * deltax
For k = 1 To 2
Worksheets(''Input'').Cells(4 + i, 10 + k).Value = U(k, i)
Next k
If U(1, i) > 0 Then
Worksheets(''Input'').Cells(4 + i, 13).Value = U(2, i) /U(1, i)
Else
Worksheets(''Input'').Cells(4 + i, 13).Value = 0
504 12 Numerical Library of Shallow Water Equations
End If
Next i
End Sub
Sub evolution()
' Program to compute the cell-averaged vector U(k,i) at the new time level using the
computed vector F at interfaces
' Flux(k,i) is the vector of fluxes at interface i + 1/2 of cell i; k make reference to
conservation laws (1 = mass, 2 = momentum)
' The domain is composed of N+2 cells; only N-2 cells are computational
' The conserved variables at cells i = 1 and N are given by the boundary conditions
dtdx = dt /deltax
For k = 1 To 2
For i = 2 To N − 1
U(k, i) = U(k, i) − dtdx * (Flux(k, i) − Flux(k, i − 1))
Next i
Next k
End Sub
Sub data()
' Program to generate the longitudinal mesh of cells
' Cell i is defined in terms of a right (index i) and left hand side nodes (index i − 1)
L = Worksheets(''Input'').Cells(4, 4).Value
deltax = Worksheets(''Input'').Cells(5, 4).Value
' number of cells
N = L /deltax
x(1) = −L /2
' coordinates of right side of each cell
For i = 2 To N
x(i) = x(i − 1) + deltax
Next i
End Sub
Sub initial()
12.2 Examples of Codes 505
' Program to read simulation data and set initial values of the dam break problem
given the upstream (Dup) and downstream (Ddown) water depths
xdam = Worksheets(''Input'').Cells(6, 4).Value
Dup = Worksheets(''Input'').Cells(7, 4).Value
Ddown = Worksheets(''Input'').Cells(8, 4).Value
CFL = Worksheets(''Input'').Cells(9, 4).Value
Tmax = Worksheets(''Input'').Cells(10, 4).Value
nM = Worksheets(''Input'').Cells(12, 4).Value
width = Worksheets(''Input'').Cells(14, 4).Value
' Computational cells are from i = 2 to N − 1; ghost cells are i = 0, 1, N and
N + 1. Cells i = 0 and N + 1 are only
' used for a fictitious reconstruction of cells i = 1 and N
For i = 0 To N + 1
U(2, i) = 0
If x(i) <= xdam Then
U(1, i) = Dup
Else: U(1, i) = Ddown
End If
Next i
End Sub
Sub MUSCL()
' Linear reconstruction of the solution U(x) within each based on the cell-averaged
values,
' with slope limiters minmod or superbee
' U(k,i) at cells i = 1 and N is required to compute Fi + 1/2 at cells i = 1 and N − 1
in the program Riemann
For i = 1 To N
For k = 1 To 2
' Computation of jumps of conserved variables upstream (U) and downstream (D)
DU = U(k, i) − U(k, i − 1)
DD = U(k, i + 1) − U(k, i)
' Limitation of the jumps in conserved variables
limiters
' MUSCL linear reconstruction of conserved variables within a cell
506 12 Numerical Library of Shallow Water Equations
Sub Riemann()
' Computation of Flux vector Fi + 1/2 at each interface using the approximate
Riemann solver HLL
' The flux Fi + 1/2 at cells i = 1 and N − 1 is required in the program evolution to
compute U(k,i) at cells i = 2 and N − 1
For i = 1 To N − 1
' Assignation of values for Riemann problem at interface i + 1/2 of cell i. CDR and
CDL are the vectors of variables U at the right (R)
' and left (L) side of the interface i + 1/2 of cell i
For k = 1 To 2
CDL(k) = BEXT(2, k, i)
CDR(k) = BEXT(1, k, i + 1)
Next k
' Transformation to physical variables flow depth (D), velocity (V), and celerity (C).
Check dry bed conditions
DL = CDL(1)
If DL = 0 Then
VL = 0
Else
VL = CDL(2) /DL
End If
CL = (9.81 * DL) ^ 0.5
DR = CDR(1)
If DR = 0 Then
VR = 0
Else
VR = CDR(2) /DR
End If
CR = (9.81 * DR) ^ 0.5
' Computation of vector fluxes FDL and FDR at interface i + 1/2 of cell i
FDL(1) = Flux1(CDL(2))
FDL(2) = Flux2(CDL(1), CDL(2))
FDR(1) = Flux1(CDR(2))
FDR(2) = Flux2(CDR(1), CDR(2))
' Compute wave estimates SL and SR for dry bed case
If DR < TOLDEP And DL > TOLDEP Then
SL = VL − CL
508 12 Numerical Library of Shallow Water Equations
SR = VL + 2 * CL
GoTo 200
End If
If DL < TOLDEP And DR > TOLDEP Then
SL = VR − 2 * CR
SR = VR + CR
GoTo 200
End If
If DR < TOLDEP And DL < TOLDEP Then
SL = −TOLDEP
SR = TOLDEP
GoTo 200
End If
DS = (0.5 * (CL + CR) + 0.25 * (VL − VR)) ^ 2 /9.81
' Compute wave estimates SL and SR for wet bed case
If DS <= DL Then
SL = VL − CL
Else
SL = VL − CL * (0.5 * DS * (DS + DL)) ^ 0.5 /DL
End If
If DS <= DR Then
SR = VR + CR
Else
SR = VR + CR * (0.5 * DS * (DS + DL)) ^ 0.5 /DR
End If
200 ' Compute the Godunov intercell flux with HLL Riemann solver
' case 1
If SL > = 0 Then
For k = 1 To 2
Flux(k, i) = FDL(k)
Next k
End If
' case 2
If SL <= 0 And SR > = 0 Then
12.2 Examples of Codes 509
For k = 1 To 2
FHLL = SR * FDL(k) − SL * FDR(k) + SL * SR * (CDR(k) − CDL(k))
Flux(k, i) = FHLL /(SR − SL)
Next k
End If
' case 3
If SR <= 0 Then
For k = 1 To 2
Flux(k, i) = FDR(k)
Next k
End If
Next i
End Sub
Sub CFLcon()
' Computation of dt to satisfy the Courant-Friedrichs-Lewy (CFL) condition and get
a stable time stepping
' Computation of the wave celerity c at each cell i using the cell-averaged values U
(1,i) = flow depth(i)
Smax = 0
For i = 1 To N
' shallow water wave celerity
c(i) = (9.81 * U(1, i)) ^ 0.5
' Local speed; where U(2,i)/U(1,i) = water velocity (i). Check zero water depths
If U(1, i) = 0 Then
SPELOC = 0
Else
SPELOC = Abs(U(2, i) /U(1, i)) + c(i)
End If
If SPELOC > = Smax Then
Smax = SPELOC
End If
510 12 Numerical Library of Shallow Water Equations
Next i
' New value of dt using the fixed value of the CFL number; i.e. 0.5 to 1
dt = CFL * deltax /Smax
' Reduce dt for early times given the approximate Smax computation
If count <= 5 Then
dt = 0.2 * dt
End If
' Check that dt+t is not greater than Tmax and recompute dt if required
If t + dt > Tmax Then
dt = Tmax − t
End If
dtdx = dt /deltax
End Sub
Public Function Flux2(U1, U2) As Double
' Computation of the flux vector component of momentum equation given vector U
If U1 = 0 Then
Flux2 = 0
Else
Flux2 = U2 ^ 2 /U1 + 0.5 * 9.81 * U1 ^ 2
End If
End Function
Public Function Flux1(U2) As Double
' Computation of the flux vector component of continuity equation given vector U
Flux1 = U2
End Function
Sub evolve()
' Computation of vector fluxes associated with PIL and PIR
FIL(1) = Flux1(PIL(2))
FIL(2) = Flux2(PIL(1), PIL(2))
FIR(1) = Flux1(PIR(2))
FIR(2) = Flux2(PIR(1), PIR(2))
12.2 Examples of Codes 511
End Sub
Sub limiters()
' Limitation of jumps in conserved variables
' If lambda = (1, Minmod limiter; 2, Superbee)
If DU * DD < 0 Then
delta = 0
Else
sig = Sgn(DU)
a = Abs(DU)
b = Abs(DD)
If a < b Then
d1 = a
Else
d1 = b
End If
d1 = lambda * d1
If a > b Then
d2 = a
Else
d2 = b
End If
If d1 < d2 Then
factor = d1
Else
factor = d2
End If
factor = factor * sig
delta = factor
End If
End Sub
Sub friction()
' Implicit computation of the friction effects
If nM = 0 Then
GoTo 250
End If
512 12 Numerical Library of Shallow Water Equations
For i = 2 To N − 1
If U(1, i) > TOLDEP Then
Rh = U(1, i) * width /(2 * U(1, i) + width)
a1 = −9.81 * dt * nM ^ 2 /Rh ^ (4 /3)
a2 = −1
a3 = U(2, i) /U(1, i)
V = (−a2 − (a2 ^ 2 – 4 * a1 * a3) ^ 0.5) /(2 * a1)
U(2, i) = V * U(1, i)
Else
U(2, i) = 0
End If
Next i
250
End Sub
Sub friction2()
' Explicit computation of the friction effects
If nM = 0 Then
GoTo 300
End If
For i = 2 To N − 1
If U(1, i) > TOLDEP Then
Rh = U(1, i) * width /(2 * U(1, i) + width)
Sf = nM ^ 2 * (U(2, i) /U(1, i)) ^ 2 /Rh ^ (4 /3)
U(2, i) = U(2, i) − 9.81 * U(1, i) * dt * Sf
Else
U(2, i) = 0
End If
Next i
300
End Sub
Sub dry()
12.2 Examples of Codes 513
t=0
count = 0
10
' Store previous solution
For k = 1 To 2
For i = 1 To N
Uold(k, i) = U(k, i)
Next i
Next k
' Compute new dt to satisfy the CFL condition
CFLcon
' Reconstruction of the solution within the cells U(x) using the cell-averaged values
U(k,i) obtained in previous time step
MUSCL
' Solve Riemann problem using the boundary extrapolated values of conserved
variables, after application of a limiter to avoid spurious oscillations
Riemann
' Evaluate bed slope source term
bedslope
' Compute cell-averaged conserved variables at new time level
evolution
' Check dry bed conditions
dry
' Include friction effects from source terms
' Implicit computation
If fric = 1 Then
friction
End If
' Explicit computation
If fric = 2 Then
friction2
516 12 Numerical Library of Shallow Water Equations
End If
' Set boundary conditions
boundary
t = t + dt
count = count + 1
' Check final of transient computations
If t <Tmax Then GoTo 10
' Print water and discharge data
For i = 1 To N
Worksheets(“Input”).Cells(4 + i, 10).Value = x(i) - 0.5 * deltax
Worksheets(“Input”).Cells(4 + i, 14).Value = z(i)
Worksheets(“Input”).Cells(4 + i, 15).Value = z(i) + U(1, i)
For k = 1 To 2
Worksheets(“Input”).Cells(4 + i, 10 + k).Value = U(k, i)
Next k
If U(1, i) > 0 Then
Worksheets(“Input”).Cells(4 + i, 13).Value = U(2, i) / U(1, i)
Else
Worksheets(“Input”).Cells(4 + i, 13).Value = 0
End If
Next i
End Sub
Sub bed()
' Definition of the bed profile
' A parabolic weir profile z=a1+a2x^* 2 is defined
a1 = 0.2
a2 = -0.05
' weir edges
xu = -(-a1 / a2) ^ 0.5
xd = (-a1 / a2) ^ 0.5
' weir profile
12.2 Examples of Codes 517
For i = 0 To N + 1
xL(i) = x(i) - 0.5 * deltax
xR(i) = x(i) + 0.5 * deltax
Next i
For i = 0 To N + 1
If xL(i) <xu Or xL(i) >xd Then
zL(i) = 0
Else
zL(i) = a1 + a2 * (xL(i)) ^ 2
End If
If xR(i) <xu Or xR(i) >xd Then
zR(i) = 0
Else
zR(i) = a1 + a2 * (xR(i)) ^ 2
End If
z(i) = 0.5 * (zL(i) + zR(i))
Next i
End Sub
Sub initial()
' Program to read simulation data
Qtarget = Worksheets(“Input”).Cells(6, 4).Value
hd = Worksheets(“Input”).Cells(8, 4).Value
CFL = Worksheets(“Input”).Cells(9, 4).Value
Tmax = Worksheets(“Input”).Cells(10, 4).Value
nM = Worksheets(“Input”).Cells(12, 4).Value
' Computational cells are from i=2 to N-1; gosh cells are i=0,1,N and N+1. Cells i=0
and N+1 are only used for a fictitious reconstruction of cells i=1 and N
'Initial condition is static water at elevation hd; dry portions of terrain allowed
For i = 0 To N + 1
If hd>= z(i) And x(i) <= 0 Then
U(1, i) = hd - z(i)
Else
518 12 Numerical Library of Shallow Water Equations
U(1, i) = 0
End If
U(2, i) = 0
Next i
End Sub
Sub boundary()
' Use ghost cells to set boundary conditions
'INLET CONDITIONS
'———————————————————
U(1, 1) = U(1, 2)
U(2, 1) = -U(2, 2)
U(1, 0) = U(1, 3)
U(2, 0) = -U(2, 3)
End If
' OUTLET CONDITIONS
'————————————————————
'Solution at cell i is vector U(k,i), with k=1 for the water depth, k=2 for the
discharge, and k=3 for the free surface elevation
U(3, i) = U(1, i) + z(i)
If U(1, i) = 0 Then
Froude(i) = 0
Else
Froude(i) = (U(2, i) ^ 2 / 9.81 / U(1, i) ^ 3) ^ 0.5
End If
Next i
For i = 1 To N
For k = 1 To 3
' Computation of jumps of conserved variables upstream (U) and downstream (D)
DU = U(k, i) - U(k, i - 1)
DD = U(k, i + 1) - U(k, i)
' Limitation of the jumps in conserved variables
limiters
' MUSCL linear reconstruction of conserved variables within a cell
PIL(k) = U(k, i) - 0.5 * delta
PIR(k) = U(k, i) + 0.5 * delta
Next k
' Values of the water depth at the interfaces of a cell for Froude<=Flim are com-
puted from SGM
If Froude(i) <= Flim Then
PIL(1) = PIL(3) - zL(i)
PIR(1) = PIR(3) - zR(i)
End If
' In a dry cell the reconstruction is not allowed; thus, water depths at interfaces are
set to zero. All the depths are zero in a cell marked computationally as dry. This is
true in a full dry cell, but incorrect in a partially-filled cell
If depth(i) <= TOLDEP Then
PIL(1) = 0
PIR(1) = 0
End If
522 12 Numerical Library of Shallow Water Equations
BEXT(2, 2, i) = 0
End If
' avoid high velocity in data
If BEXT(1, 1, i) < TOLDEP Then
BEXT(1, 2, i) = 0
End If
If BEXT(2, 1, i) < TOLDEP Then
BEXT(2, 2, i) = 0
End If
Next i
' At the interface of a full wet cell and a dry cell a jump in water depths may be
formed. However, the dry cell is in reality partially filled with water. That is, the dry
front is contained inside the computationally dry cell. Therefore, an unrealistic
dry-bed dam break wave problem at the interface of a full-wet and full-dry cells is
formed. Remember that a cell is marked computationally as dry with zero water
depth (cell-averaged value), but physically there is a portion of water inside this cell
if it is in contact with a full-wet cell
For i = 1 To N
'A reflective condition is imposed in the dry front if the water depth elevation in the
wet cell is below the average bed elevation in the (full) dry cell
If Froude(i) <= Flim Then
' dry front is on the right
If depth(i) < z(i + 1) - z(i) And depth(i + 1) <= TOLDEP Then
BEXT(1, 1, i + 1) = BEXT(2, 1, i)
BEXT(1, 2, i + 1) = -BEXT(2, 2, i)
End If
' dry front is on the left
If depth(i + 1) < z(i) - z(i + 1) And depth(i) <= TOLDEP Then
BEXT(2, 1, i) = BEXT(1, 1, i + 1)
BEXT(2, 2, i) = -BEXT(1, 2, i + 1)
End If
End If
Next i
524 12 Numerical Library of Shallow Water Equations
End Sub
Sub Riemann()
' Computation of Flux vector Fi+1/2 at each interface using the approximate
Riemann solver HLL
' The flux Fi+1/2 at cells i = 1 and N - 1 is required in the program evolution to
compute U(k,i) at cells i = 2 and N - 1
For i = 1 To N - 1
' Assignation of values for Riemann problem at interface i+1/2 of cell i. CDR and
CDL are the vectors of variables U at the right (R) and left (L) side of the interface
i+1/2 of cell i
For k = 1 To 2
CDL(k) = BEXT(2, k, i)
CDR(k) = BEXT(1, k, i + 1)
Next k
' Transformation to physical variables flow depth (D), velocity (V), and celerity (C).
Check dry bed conditions
DL = CDL(1)
If DL = 0 Then
VL = 0
Else
VL = CDL(2) / DL
End If
CL = (9.81 * DL) ^ 0.5
DR = CDR(1)
If DR = 0 Then
VR = 0
Else
VR = CDR(2) / DR
End If
CR = (9.81 * DR) ^ 0.5
' Computation of vector fluxes FDL and FDR at interface i+1/2 of cell i
FDL(1) = Flux1(CDL(2))
FDL(2) = Flux2(CDL(1), CDL(2))
FDR(1) = Flux1(CDR(2))
FDR(2) = Flux2(CDR(1), CDR(2))
' Compute wave estimates SL and SR for dry bed case
12.2 Examples of Codes 525
' case 2
If SL <= 0 And SR >= 0 Then
For k = 1 To 2
FHLL = SR * FDL(k) - SL * FDR(k) + SL * SR * (CDR(k) - CDL(k))
Flux(k, i) = FHLL / (SR - SL)
Next k
End If
' case 3
If SR <= 0 Then
For k = 1 To 2
Flux(k, i) = FDR(k)
Next k
End If
Next i
End Sub
Sub CFLcon()
' Computation of dt to satisfy the Courant-Friedrichs-Lewy(CFL) condition and get
a stable time stepping
' Computation of the wave celerity c at each cell i using the cell-averaged values
U(1,i)=flow depth(i)
Smax = 0
For i = 1 To N
' shallow water wave celerity
c(i) = (9.81 * U(1, i)) ^ 0.5
' Local speed; where U(2,i)/U(1,i)=water velocity (i). Check zero water depths
If U(1, i) = 0 Then
SPELOC = 0
Else
SPELOC = Abs(U(2, i) / U(1, i)) + c(i)
End If
If SPELOC >= Smax Then
12.2 Examples of Codes 527
Smax = SPELOC
End If
Next i
' New value of dt using the fixed value of the CFL number; i.e. 0.5 to 1
dt = CFL * deltax / Smax
' Reduce dt for early times given the approximate Smax computation
If count <= 5 Then
dt = 0.2 * dt
End If
' Check that dt+t is not greater than Tmax and recompute dt if required
If t + dt>Tmax Then
dt = Tmax - t
End If
dtdx = dt / deltax
End Sub
Public Function Flux2(U1, U2) As Double
' Computation of the flux vector component of momentum equation given vector U
If U1 = 0 Then
Flux2 = 0
Else
Flux2 = U2 ^ 2 / U1 + 0.5 * 9.81 * U1 ^ 2
End If
End Function
Public Function Flux1(U2) As Double
' Computation of the flux vector component of continuity equation given vector U
Flux1 = U2
End Function
Sub evolve()
' Computation of vector fluxes associated with PIL and PIR
FIL(1) = Flux1(PIL(2))
FIL(2) = Flux2(PIL(1), PIL(2))
528 12 Numerical Library of Shallow Water Equations
FIR(1) = Flux1(PIR(2))
FIR(2) = Flux2(PIR(1), PIR(2))
' bed source term effect
' z-jump in the bed profile within a cell
' cell full filled with water
zjump(i) = zR(i) - zL(i)
' cell partially-filled with water, with zero water depth on its right side
If PIR(1) <= TOLDEP Then
zjump(i) = PIL(1)
End If
' cell partially-filled with water, with zero water depth on its left side
If PIL(1) <= TOLDEP Then
zjump(i) = -PIR(1)
End If
deltasource(1) = 0
deltasource(2) = -9.81 * (PIL(1) + PIR(1)) * 0.5 * zjump(i) / deltax
End Sub
Sub limiters()
' Limitation of jumps in conserved variables
' If lambda=(1, Minmod limiter; 2, Superbee)
If DU * DD < 0 Then
delta = 0
Else
sig = Sgn(DU)
a = Abs(DU)
b = Abs(DD)
If a < b Then
d1 = a
Else
d1 = b
End If
d1 = lambda * d1
If a > b Then
12.2 Examples of Codes 529
d2 = a
Else
d2 = b
End If
If d1 < d2 Then
factor = d1
Else
factor = d2
End If
factor = factor * sig
delta = factor
End If
End Sub
Sub friction()
' Implicit computation of the friction effects
If nM = 0 Then
GoTo 250
End If
For i = 2 To N - 1
If U(1, i) > TOLDEP Then
If U(2, i) > 0 Then
sn = 1
Else
sn = -1
End If
a1 = -9.81 * dt * nM ^ 2 / U(1, i) ^ (4 / 3) * sn
a2 = -1
a3 = U(2, i) / U(1, i)
V = (-a2 - (a2 ^ 2 - 4 * a1 * a3) ^ 0.5) / (2 * a1)
U(2, i) = V * U(1, i)
Else
U(2, i) = 0
End If
530 12 Numerical Library of Shallow Water Equations
Next i
250
End Sub
Sub friction2()
' Explicit computation of the friction effects
If nM = 0 Then
GoTo 300
End If
For i = 2 To N - 1
' wet bed
If U(1, i) > TOLDEP Then
qadv(i) = U(2, i)
Sf = nM ^ 2 * (U(2, i) / U(1, i)) * Abs((U(2, i) / U(1, i))) / U(1, i) ^ (4 / 3)
U(2, i) = U(2, i) - 9.81 * U(1, i) * dt * Sf
' Friction can stop the flow, but a flow reversal is not physically feasible
If qadv(i) * U(2, i) < 0 Then U(2, i) = 0
Else
U(2, i) = 0
End If
Next i
300
End Sub
Sub dry()
' check dry bed conditions
' preserve positivity in water depths
For i = 2 To N - 1
' if the water depth is negative, both depth and discharge are set to zero
If U(1, i) < 0 Then
U(1, i) = 0
U(2, i) = 0
12.2 Examples of Codes 531
End If
' Notethat possible mass errors due to wet-dry updating are not corrected
Next i
' avoid high velocity in data
' if the water depth is below the tolerance (and positive) only the continuity equation
is solved, that is, discharge is set to zero
For i = 2 To N - 1
If U(1, i) < TOLDEP Then
U(2, i) = 0
End If
Next i
End Sub
Sub bedslope()
' Evaluation of source term from bed profile
For i = 1 To N
S(1, i) = 0
' Computation of the bed z-jump within a cell
' cell full-filled with water
zjump(i) = zR(i) - zL(i)
' cell partially-filled with water, with zero water depth on its right side
If BEXT(2, 1, i) <= TOLDEP Then
zjump(i) = BEXT(1, 1, i)
End If
' cell partially-filled with water, with zero water depth on its left side
If BEXT(1, 1, i) <= TOLDEP Then
zjump(i) = -BEXT(2, 1, i)
End If
S(2, i) = -9.81 * (BEXT(1, 1, i) + BEXT(2, 1, i)) * 0.5 * zjump(i) / deltax
Next i
End Sub
532 12 Numerical Library of Shallow Water Equations
The methods selected for the library are neither the only possible choices, nor uni-
versally valid. However, they are considered illustrative to introduce students to the
computation of Open Channel Flows. In this section we will describe the use of the
Library taking as example a code for flooding of an obstacle, printed in the former
section, and inserted in the file “weir_dry_terrain_SGM-DGM_MC.xls”. This solver
uses the MUSCL-Hancock scheme with a well-balanced discretization of the
bed-slope source term, along with an approximate wetting-drying procedure. The
code permits to simulate the flooding of a parabolic weir of bed profile zb = 0.2 −
0.05x2. The upstream weir side is initially wet with a water surface elevation ho,
whereas the tailwater portion is dry. We will use this code as example to show how to
use the codes available as supporting material of the textbook. Observe in Fig. 12.1
the sheet used to insert the input data, print and plot the computational results. Note
the input data selected for the current simulation.
The following input data must be supplied to the sheet for the computational
simulation:
L = length of computational domain
Δx = cell width
Upstream section:
1 = discharge pulse at inlet section fixed (qo), water depth computed by
using a backward C− characteristic
2 = discharge and water depth computed using a transmissive condition
(open boundary)
3 = discharge and water depth computed using a reflective condition (solid
wall)
Fig. 12.1 Sheet for introducing input data, printing and plotting computational results
12.3 Using the Library 533
Downstream section:
1 = discharge and water depth computed using a transmissive condition
(open boundary)
2 = discharge and water depth computed using a reflective condition (solid
wall)
ho = initial water depth at the upstream weir face. It can be lower than the
maximum weir height (0.2 m), but also higher (to simulate dam-break like flows)
CFL = Courant-Friedrichs-Lewy number
Tmax = simulation time
Limiter:
1 = minmod
2 = superbee
nM = Manning’s roughness coefficient
Friction computation:
1 = implicit
2 = explicit
Froude number for the combined Surface Gradient Method (SGM)-Depth
Gradient Method (DGM). The reconstruction of the solution is done with SGM
below this limiting Froude number, and using the DGM in other cases.
To run the model press “RUN FINITE-VOLUME SOLVER”, and the macros
are called to conduct the simulation. Once finished, numerical results are auto-
matically printed and plotted. Running the model for Tmax = 0, we observe the
initial conditions (Fig. 12.2).
In this case we fix a constant discharge pulse at the inlet section and consider the
tailwater section as an open boundary. Running the model for Tmax = 1 s, we
observe a surge approaching the weir crest (Fig. 12.3).
For Tmax = 2 s, the bore is not yet at the weir crest (Fig. 12.4); note the
deformation of the flow profile due to topography effects.
For Tmax = 2.5 s, a dry front is propagated on the lee side of the weir (Fig. 12.5).
For Tmax = 3 s, the dry front is propagated along the tailwater channel
(Fig. 12.6).
For Tmax = 5 s, the tailwater channel is wet (Fig. 12.7).
Fig. 12.5 Dry front propagation of the lee side of the weir for Tmax = 2.5 s
Fig. 12.7 Wet conditions along the entire computational domain for Tmax = 5 s
Consider now the flooding if the upstream weir side is partially submerged, e.g.
ho = 0.1 m; see simulation below for Tmax = 1 s with a surge approaching the weir
(Fig. 12.10).
During the run-up, a dry front is formed propagating in the downstream direc-
tion, while a surge propagates in the upstream direction; see simulation below for
Tmax = 2 s (Fig. 12.11).
12.3 Using the Library 537
For Tmax = 3 s, the upstream surge is still propagating, while the dry front
reaches the tailwater channel (Fig. 12.12).
The reader can access and edit the codes at the Vbasic editor (Fig. 12.13).
In the following sections examples on how to use the Library for homework
assignment to students are presented.
538 12 Numerical Library of Shallow Water Equations
subcritical flow profile, one boundary condition is prescribed at the upstream sec-
tion (typically the discharge), and another at the downstream section (typically the
flow depth). For a supercritical profile, both boundary conditions are prescribed at
the upstream section. The modification to the corresponding routine is seen in
(Fig. 12.15).
Starting with the mesh at x = 0, Fig. 12.16 contains a computation of an S2
steady gradually-varied flow profile (see input data in the figure). The corre-
sponding code is available at “backwater_MUSCLHancock.xls”.
Fig. 12.14 Modification of bed profile for computation of gradually-varied flow profiles
540 12 Numerical Library of Shallow Water Equations
Fig. 12.15 Modification of boundary conditions for computation of gradually-varied flow profiles
12.3 Using the Library 541
In Step 3, students are asked to critically look at the correctness of the steady
flow results produced by the unsteady flow solver. They are invited to compare with
a steady flow solver based on the 4th-order Runge-Kutta method, presented in
Chap. 3. The steady flow solver is available at “steadyrk4.xls”, and the results of
both solvers are compared in Fig. 12.17, showing excellent agreement. This
demonstrates the ability of the MUSCL-Hancock solver to converge to steady flow.
Note that the unsteady program correctly approaches steady uniform flow, thus the
discretization used for the friction source term is adequate.
Fig. 12.17 Comparison of steady and unsteady flow solvers for S2 steady gradually-varied flow
profile
542 12 Numerical Library of Shallow Water Equations
Wetting and drying processes over uneven topography count among the most
important problems in Environmental Hydraulics, including dam-break flows,
run-up and overland flows produced by tsunamis, flooding in rivers, and
flooding-draining of wetlands. Students are here requested to modify the code
“weir_dry_terrain_SGM-DGM_MC.xls” to simulate wetting and drying processes
over a bed profile of shape (units in m):
h i
zb ¼ 0:7 expð0:1x2 Þ þ 0:3 exp ðx 5Þ2 ; 7:9 x 7:9: ð12:1Þ
Fig. 12.18 Modification of bed topography for simulation of wetting and drying processes
12.3 Using the Library 543
At Tmax = 6 s the reflected flow at the wall produces a moving hydraulic jump
travelling in the upstream direction (Fig. 12.20).
At Tmax = 12 s the moving hydraulic jump is in its run-up over the lee-side
undulation, which is covered with a shallow supercritical flow (Fig. 12.21).
At Tmax = 18 s the moving hydraulic jump passes the lee-side undulation crest,
and the resulting wave suffers a modification due to topographic effects
(Fig. 12.22).
At Tmax = 21 s, the front of the moving hydraulic jump is conducting a run-up
over the sloping weir side, but at its tailwater portion a second shock front is formed
past the lee-side undulation crest. The flow pattern is complex but extremely
Fig. 12.23 Continuous flow profile over the lee-side hump and two-shock wave pattern with
run-up over the sloping terrain at Tmax = 21 s
In this task, students are asked to go a step further and not only modify the bed
profile, as also the initial and boundary conditions in the code, but actively con-
tribute to add any new numerical component. To produce these skills, it is requested
to investigate different discretizations of the friction source term. In Chap. 9 is
presented the explicit discretization of the friction term as [see Eq. (9.107)]
Uiadv Uiadv
qki þ 1 ¼ qadv
i ðghSf Þadv
i Dt ¼ qadv
i gn Dt
2
; ð12:2Þ
ðhki þ 1 Þ1=3
where all the variables have the meaning as explained in Chap. 9. As previously
stated, this solver works reasonably well for wet-bed conditions, but it may produce
instabilities in propagation over dry terrain if friction is not limited [see Eq. (9.110)],
given that its effect, at the limit, is to stop the flow. Thus, a flow reversal shall not be
numerically permitted.
The implicit version of the model presented in Chap. 9 is [see Eq. (9.109)]
ðUik þ 1 Þ2
qki þ 1 ¼ qadv
i ðghSf Þki þ 1 Dt ¼ qadv
i sgðqadv
i Þgn Dt
2
; ð12:3Þ
ðhki þ 1 Þ1=3
i Þgn Dt
sgðqadv 2
qadv
ðUik þ 1 Þ2 þ Uik þ 1 i
¼ 0; ð12:4Þ
ðhki þ 1 Þ4=3 hki þ 1
where the sign of the advection estimation of the flow is used to get a physically
correct solution. Implicit and explicit solvers are available in the programs used by
students during the practical sessions. Now, the task is to produce and implement a
new option for the code allowing for a semi-implicit treatment of friction.
A possible semi-implicit treatment is given by the statement
Uik þ 1 Uiadv
qki þ 1 ¼ qadv
i gn Dt
2
; ð12:5Þ
ðhki þ 1 Þ1=3
Uiadv
Uik þ 1 ¼ : ð12:6Þ
jUiadv j
1 þ gn2 Dt
ðhik þ 1 Þ4=3
Therefore, the new routine requested to the students based on their analytical
development is as follows (Fig. 12.26).
The resulting code is available in “frictioncomparison.xls”, which was tailored to
simulate dam break waves over horizontal beds. A first run was conducted using
Eq. (12.2) without limiting the maximum value of the friction force (Fig. 12.27),
thereby resulting in instabilities of the wet-dry front due to flow reversal (q < 0).
Therefore, computations are aborted shortly after this numerical artifact. This test
reveals the importance of limiting the friction force in explicit computations.
Implicit (Fig. 12.28) and semi-implicit (Fig. 12.29) simulations were both stable
and similar to each other.
12.3 Using the Library 549
The purpose of this homework is to make the students critically assess the behavior
of a code, without assuming that it works simply because computations are not
crashed. Specifically, they are asked to demonstrate that the implementation of the
reflective boundary condition is producing good physical results. At a solid wall,
U and F are given by
h h Uh 0
U¼ ¼ ; F¼ ¼ 2 : ð12:7Þ
Uh 0 U 2 h þ 12 gh2 1
2 gh
qkNþ 1 ¼ qkN1
þ1
;
qkNþþ11 ¼ qkN2
þ1
;
ð12:8Þ
hkNþ 1 ¼ hkN1
þ1
;
hkNþþ11 ¼ hkN2
þ1
;
U1 h1 U2 h2
Vw ¼ ; ð12:12Þ
h 1 h2
in Vw = −1.277 m/s and h2 = 0.3679 m. The analytical solution was easily con-
structed with the solver previously presented in Chap. 7 (Surgeanalytical.xls). The
solution of the MUSCL-Hancock solver is compared with the analytical solution at
the approaching stage to the wall (Tmax = 3.5 s, Fig. 12.30), and after reflection, at
Tmax = 7.453 s (Fig. 12.31). The agreement of the numerical and analytical results
is excellent before and after reflection, thereby confirming that the solid walls are
correctly modeled in the numerical solver.
Fig. 12.30 Numerical prediction of surge approaching the wall at Tmax = 3.5 s
hðxÞ ¼ ho expðaxÞ;
ð12:15Þ
UðxÞ ¼ Uo expð þ axÞ:
Here a is a coefficient controlling the shape of the profile and “o” refers to the
approach flow conditions. The parameter can be determined by selecting a target
flow depth hd at distance L from the inlet section, thus resulting in
1 hd
a ¼ ln : ð12:16Þ
L ho
Note that the discharge produced by the solution given by Eqs. (12.15) is
as required in steady flow without sources or sinks of mass. The Froude number is
for Eq. (12.15)
U Uo 3
F¼ ¼ exp ax ; ð12:18Þ
ðghÞ1=2 ðgho Þ1=2 2
which easily reveals the position of the critical point of the free surface profile from
2
F ¼ 1 ) xc ¼ lnðF1
o Þ: ð12:19Þ
3a
d 1 dzb
ðU 2 h þ gh2 Þ ¼ gh : ð12:20Þ
dx 2 dx
1 1
M ¼ U 2 h þ gh2 ¼ Uo2 ho ½expðaxÞ þ gh2o ½expð2axÞ: ð12:21Þ
2 2
Inserting this result into Eq. (12.20), the ODE describing the weir bed profile is
dzb U2
¼ o a expð2axÞ þ aho expðaxÞ: ð12:22Þ
dx g
This is the bed profile compatible with the analytical solution given by
Eq. (12.15) for weir flow. Finally, the free surface position is
Uo2 Uo2
zs ðxÞ ¼ hðxÞ þ zb ðxÞ ¼ ho þ expð2axÞ: ð12:24Þ
2g 2g
dzb
F¼1) ¼ 0: ð12:26Þ
dx
(a)
(b)
Fig. 12.32 Analytical (red line) and numerical (blue points) solutions for weir flow: a free surface
position b fluid velocity
References
Fenton, J.D. (2010). Numerical methods. Lecture notes, Vienna, Austria: Vienna University of
Technology.
Toro, E. F. (2000). HYPER_WAT: Library of 13 source codes for solving the non-linear
time-dependent shallow water equations in one and two space dimensions. Cheadle:
NUMERITEK Ltd.
Toro, E. F. (2001). Shock-capturing methods for free-surface shallow flows. New York: Wiley.
Author Index
Cunge, J. A., 29, 30, 201–203, 205, 210, Hancock, S. L., 383
214–216, 238–240, 278, 288, 365, 376 Harbaugh, T. E., 161, 166, 167
Harleman, D. R. F., 111
D Harrison, A. J. M., 59, 119, 121
Dawson, J. H., 116 Harten, A., 355, 361
Della Morte, R., 332 Hasumi, M., 176, 455
Denlinger, R. P., 8, 35, 39 Henderson, F. M., 5, 20, 38, 55, 59, 60, 67, 75,
De Saint-Venant, A. B., 5, 202, 211 77, 94, 105, 117, 118, 130, 140, 147,
Dey, S., 41, 130, 132, 430, 433, 461 184, 186, 187, 205, 235, 237–239, 253,
Dias, F., 462 254, 260, 285, 313, 472
Dodd, N., 369 Henry, H. R., 132
Dracos, T., 187, 188 Hicks, F. E., 216
Dressler, R., 268, 396 Hirsch, C., 290, 292, 293, 296, 367
Dressler, R. F., 7, 257, 260 Hoffman, J. D., 73, 74, 134, 144, 264,
273–276, 290, 292, 305, 319, 367, 372,
E 449, 471
Erduran, K. S., 473 Holly, F. M., 29, 30, 201–203, 205, 216, 278,
Escoffier, F. F., 187 288, 376
Evers, F. M., 410 Hosking, R. J., 196, 198, 403, 404, 446, 447,
451–453, 475
F Hosoda, T., 447, 448
Favre, H., 205, 305, 308, 396, 400 Hseng, M., 295
Fennema, R. J., 293 Hutter, K., 7, 37, 438, 458–460
Fenton, J. D., 141, 142, 244–247, 447, 493
Flannery, B. P., 134 I
Fornberg, B., 470 Iervolino, M., 427
Fraccarollo, L., 423, 427, 429 Ilic, S., 473
Fread, D. L., 161, 166, 167 Ingram, D. M., 385, 386, 388, 402, 468
Friedrichs, K., 277, 283 Ippen, A. T., 109–111, 113, 114, 116
Isaacson, E., 277, 283
G Iverson, R. M., 8–10, 17, 35, 39
Garcia-Navarro, P., 286, 288, 294, 307–309 Iwasa, Y., 187
Gardner, C. S., 461
Gersten, K., 7 J
Gharangik, A. M., 198, 199, 287, 293, Jaeger, C., 6, 30, 51, 52, 54, 59, 66, 94, 142,
399–401 165, 438
Giraldez, J. V., 37, 63, 66, 458, 459 Jain, S. C., 5–7, 25, 30, 51, 53–55, 65, 77, 108,
Glaister, P., 368 110, 117, 141, 147, 151, 156, 161, 179,
Godunov, S. K., 313, 355, 358, 378 180, 192, 202, 205, 209, 215, 217, 218,
Gottlieb, S., 373, 473 231, 232, 237, 253–255, 257, 263, 285,
Greco, M., 427 286, 313, 317, 332, 335, 337, 341, 344,
Greco, V., 334, 336, 340–345 390
Green, A. E., 439, 461 Jameson, A., 293, 300
Grilli, S. T., 467 Jan, C.-D., 165
Guinot, V., 313, 461 Jeppson, R., 55, 73, 81, 97, 102, 105, 132, 144,
313, 344
H Jia, Y., 421, 423
Hafsteinsson, H. J., 410 Jin, Y. C., 8, 19, 33, 34, 39, 435
Hager, W. H., 3–5, 10, 21, 23, 34, 37, 38,
60–64, 76, 93, 109, 116, 118, 119, 121, K
122, 124, 126, 137, 184, 186, 187, 189, Karni, S., 347
241, 247, 268, 410, 427, 433, 435, Katopodes, N. D., 19, 32, 171, 216, 240, 313,
438–443, 445–448, 454, 455, 458, 459, 344
464, 465 Keulegan, G. H., 6, 28
Author Index 559
H O
Hancock step, 383 ODE solver, 372
High-order scheme, 473 One-dimensional Serre equations, 461
HLL Riemann solver, 361 Open channel flow, 5
Horizontal plane, 19 Open channel flow classification, 1
Hydraulic jump, 98, 399
Hydraulic jump beyond a sluice gate, 198 P
Hydraulic jump position, 193 Partial gate closure, 338
Hydrologic routing, 242 Partial gate opening, 333
Hydrostatic flow, 3 Picard iteration, 438
Positive surge, 203
I Positive surge moving upstream, 210
Ideal dam break wave, 253, 392 Positive surge with friction, 308, 396
Implicit integral method, 166 Possible wave patterns, 321
Initial and boundary conditions, 221 Predictor–corrector method, 172
Initial condition, 285, 392
R
K Rarefaction wave, 230
Kinematic boundary condition, 10 Reservoir routing, 242
Kinematic wave, 234 Reynolds’ transport theorem, 205
Riemann problem, 313
L Riemann solver, 315
Lax diffusive scheme, 290 Roe Riemann solver, 368
Lax numerical flux, 365
Limitations of SWE, 241 S
Location of critical flow, 59 Saint-Venant theory, 19
Sediment transport, 41, 417
M Sediment transport closure, 421
MacCormack predictor–corrector scheme, 291 Sediment transport layer, 419
MacCormack scheme with calibrated artificial Sequent depths for general cross sections, 101
viscosity, 292 Shallow water equations, 201, 417
Method of characteristics, 217 Shock-capturing method, 214
Momentum equation, 13, 132 Shock-capturing scheme, 278
Subject Index 563