Anda di halaman 1dari 130

Computational Hydraulics

Version 2006-1

© 2005 Adri Verwey

Lecture notes

WL | delft hydraulics
Computational Hydraulics
Version 2006-1

Copyright © 2005, Adri Verwey

The right of Adri Verwey to be identified as the author of this work has been asserted in accordance
with the Copyright, Designs and Patents Act 1988

Lecture notes
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Contents

1 Introduction...............................................................................................................7

1.1 Introductory remarks......................................................................................7

1.2 Unsteady flow hydraulics ..............................................................................8

1.3 Numerical methods......................................................................................11

1.4 Mathematical modelling model ...................................................................11

2 Introduction to numerical methods.......................................................................17

2.1 Introduction..................................................................................................17

2.2 Introduction to finite difference approximations .........................................18

2.3 The Euler scheme ........................................................................................18

2.4 The improved Euler scheme ........................................................................21

2.5 The implicit scheme.....................................................................................22

2.6 The Newton-Raphson scheme .....................................................................22

2.7 A more formal analysis of accuracy ............................................................23

2.8 Stability........................................................................................................26

2.9 Consistency and convergence ......................................................................27

2.10 The choice of a time step .............................................................................27

2.11 Reservoir routing .........................................................................................29

2.12 Routing of a decayable pollutant through a reservoir..................................33

2.13 Non-uniform steady flow in channels..........................................................34

2.14 An inverse scheme for the backwater curve computation ...........................38

2.15 Non-uniform flow in non-uniform channels................................................40

2.16 What have we learnt?...................................................................................41

2.17 Questions and small assignments ................................................................41

WL | Delft Hydraulics 3
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

3 Basic Unsteady Channel Flow Equations .............................................................43

3.1 Introduction..................................................................................................43

3.2 Continuity equation .....................................................................................43

3.3 Momentum equation....................................................................................44

3.4 Transformation to the characteristic form....................................................45

3.5 The significance of the characteristics.........................................................48

3.6 The method of characteristics ......................................................................51

3.7 More complex boundary conditions ............................................................55

3.8 The formation of positive and negative hydraulic jumps ............................57

3.9 The limited practical importance of the method of characteristics..............59

3.10 Questions and assignments ..........................................................................60

4 Introducing Numerical Solutions for Partial Differential Equations.................61

4.1 Introduction..................................................................................................61

4.2 The advection equation................................................................................63

4.3 The characteristic solution ...........................................................................64

4.4 Finite difference schemes ............................................................................65

4.5 Characteristic solutions on a fixed grid .......................................................70

4.6 Introducing diffusion ...................................................................................72

4.7 An explicit finite difference scheme ............................................................73

4.8 Explicit schemes for the combination of advection and diffusion...............76

4.9 Implicit schemes for the diffusion equation.................................................77

5 De Saint Venant equations and their solutions.....................................................81

5.1 Introduction..................................................................................................81

5.2 The continuity equation ...............................................................................82

5.3 The momentum equation .............................................................................84

5.4 Numerical solutions .....................................................................................88

WL | Delft Hydraulics 4
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

5.5 Description of hydraulic structures..............................................................92

5.6 Topological model schematisation...............................................................94

5.7 Hydraulic model schematisation..................................................................95

5.8 Boundary- and initial conditions..................................................................97

5.9 Model calibration and validation .................................................................99

6 Mathematical modelling of floods .......................................................................101

6.1 Introduction................................................................................................101

6.2 Flood model requirements .........................................................................101

6.3 The role of new data collection technologies ............................................103

6.4 The nature of flood wave propagation .......................................................104

6.5 Deformation of flood waves ......................................................................105

6.5.1 The role of varying celerities ........................................................105

6.5.2 The role of the diffusion term .......................................................106

6.5.3 The role of the lateral flow terms..................................................107

6.6 Link to hydrologic flood routing models ...................................................107

6.7 Two-dimensional modelling of floods .......................................................108

6.8 Integrated 1D/2D modelling ......................................................................113

6.9 Exercise......................................................................................................116

7 Water Hammer .....................................................................................................117

7.1 Introduction................................................................................................117

7.2 Water Hammer Equations ..........................................................................119

7.3 The Method of Characteristics for Water Hammer....................................125

7.4 Exercise......................................................................................................128

8 References..............................................................................................................129

WL | Delft Hydraulics 5
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

WL | Delft Hydraulics 6
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

1 Introduction

1.1 Introductory remarks


The topic of computational hydraulics is dealing with the question how water
resources engineers and planners can be assisted in dealing with complex hydraulic
problems. When these problems are very local, they can generally be addressed by
using empirical relationships and a well trained engineer usually is equipped with
methods and tools dealing with it. In these cases, the use of laws and relationships in
hydraulics is often limited to steady flow approximations.

For larger scale problems, however, the unsteady nature of flows becomes more
dominant and methods used will become more complex. Whereas until various
decades ago, the focus has been on developing approximations, partly empirical, of
the full hydrodynamic behaviour of the water system, the use of computers has made
it possible to describe hydraulic systems quite accurately. Over the past decades
enormous progress has been made in developing simulators or mathematical models
for all kinds of hydraulic systems, such as rivers, drainage networks, irrigation
networks, water distribution networks etc.

Currently, the level of accuracy of such simulators is primarily limited by the quality
of data available to construct and calibrate the models. A variety of good software
packages is available to construct such models. However, good schematisations of
hydraulic systems in models requires some insight in the laws and techniques behind
these modelling systems in order to use them correctly in building one’s own model.

In this series of lectures we address this need by providing insight into the nature of
one-dimensional unsteady flow. After the introduction of the unsteady flow
equations, the link between the equations and the physical system is shown by the
characteristic celerities of disturbances propagating along channels. This provides
the basis for the numerical schemes developed to solve the equations. Moreover, it
shows clearly the effect of boundary variations and, in particular, the effects of
control of hydraulic systems.

With this understanding as a basis, numerical method is introduced. First, only the
so-called ordinary differential equations are treated, enabling us to do simple
backwater computations, water quality simulation in well mixed reservoirs etc.
However, at the same time numerical concepts and their evaluation are introduced
relating to the accuracy, stability, robustness and efficiency of numerical operations.
This will serve as a necessary basis for those who want to develop their own models,
as well as for those applying existing modelling systems.

WL | Delft Hydraulics 7
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Subsequently, numerical descriptions of more complex problems described by


partial differential equations are discussed, primarily to provide enough insight as a
basis for good mathematical modelling practice. Also here, the role and choice of
numerical parameters in mathematical model simulation is emphasised.

Finally, the overall construction of good mathematical models is discussed. Based


upon a clear idea about the objective of model use, rules for the most economical
and correct schematisation are developed, starting with a topological schematisation
of the system. In addition, the derivation of correct physical parameters is discussed,
such as those describing channel conveyance, storage and lateral flows. Moreover,
the possibilities and limitations of such models is discussed, with special attention
given to the use of models under extrapolated conditions, such as often the case in
the modelling of floods.

1.2 Unsteady flow hydraulics


In teaching, the topic of hydraulics is generally first introduced from the steady flow
concept. In principle, it would be more logical to start with the unsteady flow
concept and than introduce steady flow as a special case of unsteady flow. In this
case a clear indication has to be made of the underlying assumptions and the
simplification of the problem. In each application these assumptions have to be
verified.

It should be realised that a real steady flow does not exist in nature. There will
always be some small variations in the flow distribution, even if there are no
observable water level variations. The steady flow is a concept of our engineering
mind, which sometimes can simplify engineering without unacceptable differences
from reality. However, the danger exists that engineers turn too easily towards the
concept of steady flow, even in cases where this is a quite wrong schematization of
reality and where this approach may lead to quite wrong conclusions.

For this reason, it is important to familiarize oneself with problems which typically
show more significant variations both in time and in space. Hydraulic problems in
channels are governed by two important concepts: storage and conveyance. For
incompressible flow, the first concept deals with water volume conservation. The
second concept deals with balance of forces acting upon the water mass and their
effect on the momentum balance. Of particular importance in this description is the
magnitude of flow momentum losses due to channel friction relative to the gain of
momentum due to gravity or other forces.

Before we can discuss more in detail the difference between steady and unsteady
flow, the concept of boundary conditions has to be introduced. In general, one is
only interested in a specific part of a hydraulic system. To illustrate this idea, it is
useful to consider the hydrological cycle. Water evaporates from the sea surface,
precipitates partly above land and flows via the land surface, or via infiltration
through the subsurface, to the rivers and, in most cases, back to the sea. Let us now
consider the river part of this cycle, or even a small part of this river system. The
link to the upstream part of this river subsystem is specified in the form of a
boundary condition and, more specifically, as the upstream boundary condition.

WL | Delft Hydraulics 8
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The link to the lower part of the river system, or to the sea, is specified as the
downstream boundary condition. In the sequel, we will discuss these boundary
conditions in more detail, including the question of the real need for such conditions
in our computations.

The question of the flow type is very much linked to the nature of the changes at the
boundaries of the area described (or modelled). When the part of the river system
studied adapts its state only slowly to the changes at the boundaries, the system is
considered to be unsteady. However, when the river system adapts its state nearly
instantaneously to the changes at the boundaries, the system is said to be quasi-
steady and some steady flow concepts might be used. This is usually the case when
local or near-field problems are studied, such as local structures with their typical
backwater effects.

One of the important parameters influencing this ease to adaptation is the storage in
the system. If this storage capacity is large compared to the difference between
inflow and outflow the time scale of adaptation is also large. It is very likely, then,
that we will observe a strongly unsteady flow phenomenon. If, however, there is
little storage capacity available between the boundaries, the adaptation of the state of
the system to the new boundary conditions may be fast and the flow may pass
through a sequence of nearly-steady states. This adaptation is also dependent on the
facility of the flow to accelerate or decelerate. If the adaptation of flow particle
velocities to changing boundary conditions is fast, the system will pass through a
series of nearly-steady states. Where the adaptation is slow compared to the changes
at the boundaries, the result will be clearly an unsteady flow.

These concepts are best illustrated by giving some practical examples.

Flood waves in rivers

Floods in a river are the result of the surface- and subsurface runoff generated during
periods of intense rainfall. This response usually leads to a typical hydrograph shape
discharge and water level variation in the river, which is more pronounced when the
rain is uniformly distributed over the period of the rain event. However, as rainfall is
generally not uniformly distributed in time, the discharge distribution from the
catchments into the river is usually less irregular. While the flood wave propagates
down the river it undergoes further deformations due to varying storage and
conveyance characteristics of the river channel and due to additional lateral inflows
from other catchments. These processes together form a typical unsteady flow
phenomenon, which can only be studied by simulations on the basis of unsteady
flow equations.

Flow in an urban drainage or combined sewer system

Drainage from urban catchments is to be seen as a special case of flood routing.


Differences with a rural catchment are the faster response to the rainfall, the smaller
amount of storage usually available and the more important role of channel
conveyance, compared to channel or reservoir storage. In systems with drainage
pipes, the open channel flow may temporarily change into a pressurized flow. The
unsteadiness of the phenomenon, however, is very pronounced.

WL | Delft Hydraulics 9
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Despite these differences, it will be shown in the sequel that these problems require
the same modelling tools as the ones used for the study of flood routing in rivers.

Flow in a desert wadi

In wadis, the unsteady nature of the flow is even more pronounced than in many
other rivers, due to the infiltration of the water into the river and flood plain bed.
This leads to the steepening up of the flood wave front. Also, if often leads to the
complete disappearance of the flood wave after some time.

Flow in irrigation systems

Flow in irrigation systems is usually controlled for the optimum use of the water
resources. This control introduces variations in time. A fast control may even lead to
the formation of hydraulic jumps travelling along the channel. Although the design
of the irrigation canals is often based on steady flow concepts, the performance
under operation usually requires checks on the basis of unsteady flow computations.

Flow over or through a hydraulic structure

The description of flow through a hydraulic structure within a river branch is usually
based on an assumption of steady flow. The discharge at each moment in time is
directly dependent on the water level boundary conditions. Although these water
levels may vary rather fast, the discharge generated will respond more or less
instantaneously. The immediate adaptation of the flow to changing boundary
conditions is the result of the lack of storage between the upstream and downstream
section and the presence of a relatively small water mass to be accelerated or
decelerated. As will be shown in the sequel, the possibility to link a steady flow
channel element to channel elements where the flow has a distinctly unsteady nature,
depends on the relative importance of the various terms of the equations describing
the flow problem.

Flow in pipe networks

Flow in pipe networks for water distribution in a town is often computed as steady
flow. For given constant water demands at various places in town and constant water
levels in reservoirs, the flow and pressure distribution over the complete network
can indeed be computed. These computed pressures can be checked against
minimum pressure requirements. However, the assumption of constant demands is
an oversimplification of reality. Water demands usually have daily and weekly
cycles. Reservoirs may be filled during the night at cheaper electricity rates and
emptied during the day, when demands are higher. Fire fighting may suddenly
change the water demands over the network. These varying demands and storage
lead to unsteady flow phenomena in pipe networks, although these are still often
computed as series of quasi-steady states. Complete unsteady flow may occur as a
result of sudden changes caused by failures or misoperation of the distribution
network. This may cause unacceptable water hammer and cavitation effects. In this
case computations are based on the use of the full unsteady flow equations for pipe
flow, including storage of water resulting from water compressibility and pipe
diameter expansion.

WL | Delft Hydraulics 10
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

1.3 Numerical methods


Although numerical methods have been developed already for centuries, the
possibilities offered by computers have given a strong impulse to the further
development of these methods. With numerical methods one tries to solve sets of
differential equations, such as the De Saint Venant equations for unsteady flow in
river- and channel systems. In the derivation of the differential equations one starts
with finite control volumes for the definition of balance equations and than assumes
that these volumes reduce to an infinitesimal small size. Under this assumption,
these equations provide a correct and valid description of continuous flows.

With numerical methods this process is reversed and balance equations are derived
over finite control volumes, starting from the differential equations. For this reason,
one cannot expect that this procedure leads to the same results as those which might
have been obtained by solving the partial differential equations directly. However, in
most cases such solutions do not exist, particularly when the equations are non-
linear. This, unfortunately, is usually the case when solving practical problems in
hydraulics.

Currently, for nearly all applications in hydraulics, numerical methods have been
developed that work well and have the potential of limiting the differences between
exact solutions and the approximate solutions. In these lectures we are discussing the
differences by dealing with concept such as consistency, stability, robustness and
economy of numerical operations. In particular, it will be shown how partial
differential equations can be transformed into linear finite difference equations, to
which extent these linearization’s require iterations and what sort of algorithms exist
to solve the systems of equations in an economical way.

It will also be shown, wherever applicable, that numerical behaviour is related to the
physical behaviour described. Most obvious is the relation between boundary data
requirements and the way changes at boundaries are affected by the hydraulic
system. However, also the performance of iteration techniques are influenced by the
physical behaviour of the system.

The objective of dealing with numerical methods in this lecture series is to provide
enough background information to serve as a basis for the correct development of
models and for the best choice of modelling systems offered for use in a project.

1.4 Mathematical modelling


Modelling has become a frequently used tool for studies in hydraulic and
environmental engineering. Whereas in the past many engineers turned to physically
based models or simplified descriptions for the support of engineering studies, the
increasing availability of personal computers and the powerful developments in
computer graphics, data bases and on-line control, software has brought computer
support to the desk of consulting engineers. In line with these developments we also
see a strongly increased availability and use of mathematical modelling software
tools.

WL | Delft Hydraulics 11
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

For a better understanding of what a model represents, let us look at one of the many
possible ways of defining such a tool:

a model is a physical or mathematical description of a physical system,


including the interaction with its outside world, which can be used to simulate
the effect of changes in the system itself or the effect of changes in the conditions
imposed upon it.

In the development of a mathematical model one may distinguish the following


main elements:

· definition of objectives
· schematisation
· equations and conditions
· solution algorithm
· software choice or development
· data collection
· model calibration
· model verification
· simulations

Definition of objectives

Definition of objectives is a very important and often underestimated element in the


decision process which leads to the use of a model. The first question to be posed is
whether a model can add important information to what is already understood about
a system. It also involves the estimation of the possibility to save project costs by
using a model and the economic value that may be attached to the development and
use of the model. This process will also lead to the choice of the level of complexity
of model description. The use of models is generally associated to their use for
simulations. However, another interesting field of application is in the analysis of
possible hypothesis about empirical relations. Finally, models define relations
between the various variables describing a state of a physical system and
consequently models may be used for data consistency checking. As typical
examples of model objectives in the fields of environmental, hydraulic and
hydrological engineering one may mention:

· effects of hydraulic works;


· simulation of the impacts of floods;
· on-line flood forecasting;
· flood prediction;
· design of urban drainage systems;
· simulation of the impacts of dam breaks;
· study of field irrigation water supply;
· control of salt intrusion in estuaries;
· BOD-DO computations along rivers;
· ecological effects of heat loads from power plants;
· estimation of sedimentation in reservoirs;

WL | Delft Hydraulics 12
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

· soil remediation studies;


· effects of sewerage overflows upon the receiving waters;
· consistency checking of water quality data;
· consistency checking of hydrological data.

Model schematization

The schematisation of the physical system follows from the complexity of the
processes and the economical interest in studying these processes in all their details.
The use of simplified models, based on a simplified description of the physical
processes is only justified if the results of the model can still be used reliably in the
design process. In other words, if the results still fall within the reliability range of
data used by the designer. All processes in nature are of a three-dimensional and
unsteady nature. The choice in the model schematisation is primarily:

· choice of the number of spatial dimensions;


· choice of time variability.

From the spatial dimensions and the time, the modeller selects one or more
independent variables x,y,z or t, or other independent variables if certain
transformations are applied, e.g. r and in a polar coordinate system. Such a choice
is strongly linked to the model objectives. For example, a reservoir can be
schematized into a single point, if one wants to study the water level variations and
the reservoir outflow as a function of time. The same reservoir, however, will
require a three-dimensional schematization in space in a study of wind-induced
circulations or velocity patterns following from density stratification.

Equations and boundary conditions

A mathematical model is based primarily on the choice of equations describing the


state of the physical system. Water levels, discharges, velocities, temperatures,
salinities etc. are so-called state variables or dependent variables. One equation is
needed for each of these variables describing the state of the physical system. Most
of these equations are balance equations of mass, or, simplified, of volumes. Other
equations are based on balances of other physical quantities, such as momentum,
energy or turbulence. Often, also, simplified forms of these equations are used for
the description of processes within our computational domain in space and time. In
space this domain might represent the axis of a river flowing from town A to town B
further downstream and in time the duration of a typical flood wave, including an
antecedent period. Additional conditions have to be provided at the model
boundaries, to specify the interaction of the outside world with the domain described
by our model. These conditions will follow from the nature of the physical
processes, translated into mathematical conditions through the equations describing
the system.

WL | Delft Hydraulics 13
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Solution algorithms

The balance equations in space and time provide us, generally, with partial
differential equations. These equations are transformed from a continuous form into
a discrete form by writing them out as relations between variables at points in the
computational domain. Such formulations are based on finite differences, finite
elements or boundary integrals and provide systems of linear equations. Completed
with a linearized formulation of boundary conditions and special elements, such as
hydraulic structures, the total system of linear equations may be solved with a
variety of algorithms, ranging from simple Gaussian elimination as a direct matrix
solver, to iterative techniques such as the conjugate gradient method. The solution
algorithm usually has to follow a specific sequence of operations, consistent with the
physical links in the system.

Software

For most environmental studies standard software tools are available. For simple
problems engineers usually turn to spread sheet packages, whereas for problems
related to open channel flow in networks, reservoirs, flow in pipe networks, sewer
systems, water hammer, coastal management, short waves computations etc. various
software products are available, developed at specialized hydraulic research
institutes. The use of these packages assures a flexible user environment and a
reliable solution of all sorts of numerical problems.

Data collection

Over the past years more and more effort has been spent on the collection of all sorts
of data and the processing and storage of these data in data bases. However, for
model development, data available in data bases is not always sufficient, as the
calibration of models often requires data measured over short periods in time,
available at various locations simultaneously. For this reason, the models are usually
set up with whatever data available in the standard data base, completed with data
collected during some specially organised campaigns.

Model calibration and verification

Some data required for a model can be collected directly in the field. Examples are
salinities, channel cross-sections, discharges, water levels and concentrations of
dissolved substances. Some data can only be collected with a certain degree of
uncertainty, such as the details of the topography in the flood plains. Other types of
data cannot be measured at all, such as Manning numbers and diffusion coefficients.
Such data can only be estimated on the basis of a sound engineering judgement,
based on the interpretation of recorded values of other variables and parameters. The
more uncertainty we have in the model parameters, the more we are dependent on a
good set of calibration data. The fit between measurements and computations and
the knowledge of the processes enables the adjustment of the parameters until an
acceptable fit has been obtained. For model calibration one will usually select a
number of events, which are complementary to each other in terms of the calibration
parameters.

WL | Delft Hydraulics 14
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

A calibration of a river model, for example, will typically start with low flow
calibration and finish up with the calibration of some typical flood events. This
procedure allows for the calibration of channel roughness parameters, prior to the
calibration of flood plain conveyance and storage parameters. After completion of a
model calibration the model should be verified on a set of data not yet used for
parameter estimation. However, in practise it is not easy to reserve such a set and
even if such verification runs are made, the differences between model and
prototype performance may lead to lengthy arguments about the quality of the model
data. In other words, why should one trust the results of verification more than the
results of a model calibration?

Simulations

Once the model has been accepted it can be used for the typical simulations
following from the definition of the model objectives. It should be kept in mind that
the use of the model with modified parameters may, in turn, modify other
parameters. As an example, the construction of river embankments may also change
the bed roughness. The use of such models, therefore, should always be
accompanied with sound engineering judgement based on a thorough knowledge of
the physical processes.

WL | Delft Hydraulics 15
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

WL | Delft Hydraulics 16
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

2 Introduction to numerical methods

2.1 Introduction
This chapter deals with the solution of problems described on the basis of one single
independent variable, such as x or t. A typical example is the simulation of flow
through a reservoir, where water level variations in the horizontal plan are neglected
and the point water level fluctuates in time as a result of time-variations in inflow
and outflow. The water volume balance leads to an ordinary differential equation of
the first order which can conveniently be solved by a finite difference
approximation.

Another typical example connected to this problem is the description of the variation
in time of the concentration of a chemical substance in the reservoir water, assuming
that this reservoir is well-mixed. The balance equation for this substance also leads
to an ordinary differential equation with time as the independent variable.

In a further extension one may consider the interaction of various substances,


leading to coupled systems of first-order ordinary differential equations. A well-
known example is given by the Streeter-Phelps equations, describing the decay of
organic pollutants and the effects on the oxygen concentration in a well-mixed pond.
The same set of equations is found by considering the BOD and DO concentrations
in a particle of water moving with the stream velocity in a river and neglecting the
influence of exchange by diffusion between various water particles. Although this
set of equations may be described as a function of time, it might also be convenient
to apply a transformation which takes the channel axis x as the independent variable.
Such transformation facilitates the inclusion of typical influences along the river,
such as point effluent loads, weirs providing additional reaeration and the effects of
variations in the channel topography on stream velocity and reaeration.

Although the simplest problems of this category can be studied by exact solutions of
the differential equations, the more flexible formulation of model equations and
parameters requires solutions formulated by numerical schemes, such as finite
difference schemes or finite element schemes. In the sequel, a range of finite
difference approximation are introduced and discussed in terms of accuracy, stability
and convergence of solutions.

Figure 2.1 Sketch of a cylindrical groundwater reservoir

WL | Delft Hydraulics 17
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

2.2 Introduction to finite difference approximations


Let us consider one of the simplest ordinary differential equations expressing the
volume balance of groundwater in a cylindrical reservoir as shown in Figure (2.1).
Assuming a constant reservoir surface area A, a constant porosity n and an outflow
defined as a linear function of the hydraulic head h, the water level is described by
the continuity equation and Darcy law, respectively, as

dh
nA = -Q ; Q = - kAh
dt

or, by elimination of Q,

dh
Article 2. = -ah (2.1)
dt

where is a linear reservoir coefficient.

This equation has the exact solution

h = h0 e-a t (2.2)

where h0 is the initial water level in the reservoir.

This exact solution is given here primarily with the purpose of comparing various
numerical schemes, solved with a variety of numerical parameters for these
schemes. The numerical schemes that will be introduced successively are the

· Euler scheme;
· Improved Euler scheme;
· Implicit scheme;
· Newton-Raphson scheme.

The numerical parameter in this example is the time step of numerical integration t.

2.3 The Euler scheme


For the definition of a finite difference approximation we will first return to the
concept of derivatives and the differential equations that may be constructed from
these. The meaning of Equation (2.1) is that at any point along t, the derivative dh/dt
to the function h(t) is equal to the value of the right hand side of that equation.

WL | Delft Hydraulics 18
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 2.2 Situation sketch for the definition of a derivative (tangent) to a function h(t)

Keeping in mind (Figure 2.2) that this derivative is defined as

dh h(t + Dt ) - h(t )
= L im (2.3)
dt Dt ® 0 Dt

one may use the inverse of this definition to construct a finite difference scheme of
Equation (2-3) on the basis of the approximation
dh Dh h n +1 - h n
@ = @ - a hn
dt Dt Dt
or
h n +1 @ (1 - aDt ) h n (2.4)

where it should be noted that n is introduced as a counter for the time step and
written as a superscript to the variable h or to any other quantity defined at a grid
point at time t=n t. This notation should not be confused with an exponent. The
numerical scheme introduced this way is called an Euler scheme. The right hand
side of Equation (2.6) is taken at time t=n t. Assuming that the value of h is known
at that point, we call such a scheme a forward difference scheme as we construct a
solution forward in time proceeding from a point where the solution is already
known.

Figure 2.3 Sketch of a finite difference approximation as the inverse of the definition of a
derivative at a point A at t=n t

WL | Delft Hydraulics 19
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure (2.3) also shows that for any non-linear solution the finite difference scheme
introduces at each time step an error , defined as the difference between the exact
solution of the equation and the solution obtained with the scheme. From the figure
it may also be concluded that this error usually increases with an increasing time
step. It will also be clear that the error depends on the curvature of the function,
expressed by the influence of the higher-order derivatives. In other words, for
solutions deviating slowly from straight lines one may take larger time steps than for
solutions which deviate fast from straight lines, if one wants to obtain the same
relative accuracy of the solution.

Let us demonstrate the effect of the choice of time step by solving Equations (2.1)
and with the Euler scheme of Equation (2.4) for the following data:

h0 = 10 mm
= 0.1 day-1

Table 2.1 shows results at T=4 days for the exact solution of the equations,
compared with numerical solutions obtained with Equation (2.4) for time steps t =
0.5 days, 1.0 day and 2.0 days respectively. Differences between the exact solution
and the numerical solution are 1 %, 2 % and 4.5 %, respectively. It is also observed
that the errors are larger than those found at T = 2 days, demonstrating the
accumulative effect of the errors during the integration of the differential equation
along the time axis.

Table 2.1 Influence of the numerical scheme and the time step on the accuracy of results

time Exact Euler Improved Euler Implicit


(days) t=½ t=1 t=2 t=1 t=2 t=1 t=2
(days) (days) (days) (days) (days) (days) (days)
0.0 10.00 10.00 10.00 10.00 10.00 10.00 10.00 10.00
0.5 9.50 9.50
1.0 9.05 9.03 9.00 9.05 9.00 9.05
1.5 8.57 8.60
2.0 8.19 8.15 8.10 8.00 8.19 8.20 8.19 8.18
2.5 7.74 7.78
3.0 7.41 7.35 7.29 7.41 7.38 7.41
3.5 6.98 7.04
4.0 6.70 6.63 6.56 6.40 6.71 6.72 6.70 6.69

WL | Delft Hydraulics 20
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 2.4 Various derivatives used in the finite difference approximation: (a) derivative used in
the Euler scheme; (b) ideal choice of derivative bringing the solution from point A to
point B; (c) approximate derivative used in the centred schemes

2.4 The improved Euler scheme


From Figure (2.4) it is observed that one possible way of reducing the errors is by
selecting a better location for the derivative along the time axis. A good location
would be a place C, approximately halfway the time levels n t and (n+1) t.
However, the exact location of the point is not known, due to the influence of the
higher-order derivatives on the shape of the solution function. Moreover, even if this
exact location would be known, the solution of the function is not known at this
point and, hence the value of the derivative. The principle of a better positioning of
the derivative leads us to the improved Euler scheme based on an additional iteration
of the solution. As a first step in the improved Euler method the value of hn+½,
halfway the time step, is approximated by

h n +½ @ (1 - ½aDt ) h n (2.5)
Substitution of this approximate value in the expression of the derivative gives the
finite difference approximation over the total time step from grid point n to (n+1) as

h n +1 - h n
@ - a h n +½
Dt
or
h n +1 @ h n - aDt h n +½ (2.6)

Results of this scheme, also given in Table (2.1), show considerable improvements
in accuracy, with errors of 0.15 % and 0.3 % for time steps of 1.0 and 2.0 days,
respectively. Even considering that the amount of computational work done with the
improved Euler scheme is approximately twice the amount of work done with the
normal Euler scheme, the improvement in terms of efficiency is still remarkable.

For an equivalent computational effort the improved Euler scheme produces only 15
% of the error of the normal Euler scheme. Although similar improvements in
accuracy are not always obtained for all problems, the example demonstrates the
potential of efficiency improvements by using higher-accurate numerical schemes.

WL | Delft Hydraulics 21
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

2.5 The implicit scheme


The numerical schemes introduced so far are based on the concept that the derivative
of the function is known at the initial point in the computation over a time step and
does not change during that time step or, at best, is adapted in its value and location
during an additional iteration. A next logical step for improvement, then, is to
include the variables composing the derivative in the expression for the yet unknown
variable h at time level (n + 1) t. For Equation (2.1) this leads to a centred implicit
finite difference scheme of the form

h n +1 - h n
@ - (½a h n + ½a h n +1 )
Dt

or
1 - ½aDt n
h n +1 = h (2.7)
1 + ½aDt

For the given value of and a time step of 1 day this leads to the simple relation

h n +1 = 0.9048 h n .

Results of this computation are shown in Table (2.1). The errors are 0.017% and
0.13% for time steps of 1.0 and 2.0 days, respectively. In terms of accuracy, this
approach gives another considerable improvement over the earlier introduced so-
called explicit schemes. Although this conclusion may not be generalized to other
numerical schemes without exceptions, the implicit schemes, in general, have the
potential of providing more accurate results, at lower computations cost, due to the
better centring of the finite difference equations. However, for problems involving
more unknown variables, the implicit schemes lead to systems of linear equations,
which have to be solved simultaneously through matrix operations.

In most cases the overall solution algorithm leads to more numerical operations per
time step. However, this is generally compensated by the much larger time steps that
can be taken. As a consequence, currently most numerical algorithms are based upon
implicit schemes. Apart from the higher accuracy, another important advantage of
implicit schemes is their improved stability or robustness behaviour, as discussed in
§ 2.8.

2.6 The Newton-Raphson scheme


One special form of implicit schemes is the Newton-Raphson scheme. Although the
Newton- Raphson approach is generally presented as a method for solving nonlinear
equations, we introduce it here as an approach to the formulation of linear finite
difference schemes and notably the linearization of individual terms and coefficients
in the scheme.

WL | Delft Hydraulics 22
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Assume, for example, that the coefficient in Equation (2.1) is given as a linear
function of h by the relation

a = a0 + a1h

where a0 and a1 are constants. Substitution into Equation (2.1) gives

dh
= - a0 h - a1h 2 = F (2.8)
dt

where F is a general function of h. An implicit finite difference scheme, centred


halfway the time step, could be written as

Dh
= F n + ½DF
Dt

As F is a function of h, the Newton-Raphson approach gives


dF
DF = Dh = - ( a0 + 2a1h ) Dh
dh

leading to the finite difference scheme

F n Dt
Dh = (2.9)
1 + ½Dt ( a0 + 2a1h n )

The advantage of this approach is that the change in the value of the coefficient is
already partly taken into account during the integration over the time step. From
Equation (2.13) one may conclude that the process is not yet ideal, as the value of h
in the right hand side of the expression is taken at grid point n, whereas the
substitution of a value at grid point (n+½) would be more precise. Referring, again,
to the implicit scheme of § 2.5, the value of would be taken at grid point n,
whereas an improved centring would require an additional iteration, similar to the
approach followed in the improved Euler method. The Newton- Raphson implicit
scheme is generally used without such additional iteration as in most cases this extra
step is hardly cost effective.

2.7 A more formal analysis of accuracy


Although the concept of accuracy was discussed on the basis of a common sense
approach and such a reasoning should always accompany the use of numerical and
physical concepts, it is also useful to introduce more fundamental analysis
techniques. One of the most frequently used tools for the analysis of the accuracy of
numerical schemes is the Taylor's series expansion.

WL | Delft Hydraulics 23
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Referring to Equation (2.2) and assuming a continuous behaviour of the function h


and all its first and higher-order derivatives in time, the value of h at grid point n+1
can be expanded from its value at grid point n through the infinite Taylor's series

n
Dt k æ d k h ö
¥
h n +1
= h +å
n
ç k ÷ (2.10)
k =1 k ! è dt ø

with all derivatives taken at grid point n. In Equation (2.10) k! should be read as “k
factorial”, defined as the product 1*2*3*… .*k.

It will be useful to discuss the meaning of the term under the summation sign in a
pragmatic way. The term tk refers to a step in time raised to the power k and
cancels, at least in magnitude, against the contribution dtk. The notation dkh has the
meaning of expressing differences in function values at points in the vicinity of the
point where the Taylor's series is expanded upon and has a value comparable in
order of magnitude to the average function value at these points. As the value of k!
increases rapidly with increasing k, it may be expected then that the contribution of
the higher-order derivative terms in this expanded series decreases rapidly with
increasing k.

Referring to Equation (2.10), one may divide all terms by t to give

n n n
h n +1 - h n æ dh ö Dt æ d h ö Dt æ d h ö
2 2 3
= ç ÷ + ç 2÷ + ç ÷ + h.o.t . (2.11)
Dt è dt ø 2 è dt ø 6 è dt 3 ø

where h.o.t. refers to all higher-order terms in this series expansion. Substitution of
the Euler scheme of Equation (2.4) into Equation (2.11) and dropping the superscript
n provides

dh Dt d 2 h Dt 2 d 3 h
= -ah - 2
- 3
+ h.o.t . (2.12)
dt 2 dt
14444244443 6 dt
-TE

Comparison of Equation (2.12)with Equation (2.1) leads to a difference TE between


the differential equation and the Euler scheme defined with the intention to solve
this equation.

This difference TE is called the truncation error of the finite difference scheme. For
the improved Euler scheme, Figure (2.3) visualizes this truncation error by the
magnitude . For the special case of Equation (2.1), the magnitude of all higher-
order derivatives can be expressed in terms of the value of h at those points, as
shown for constant by successive differentiation of the equation with respect to t.

WL | Delft Hydraulics 24
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

At all points in time this gives the truncation error in the form

a2 h a3 h 2
TE = Dt - Dt + h.o.t. (2.13)
2 6

From Equation (2.13) it is readily seen that by decreasing t in the integration of


Equation (2.1), the total error in h decreases linearly with t with respect to the first
term in the right hand side of Equation (2.13) and decreases even faster than linearly
with respect to the remaining terms in the truncation error. A truncation error of this
type is said to be of first order in t or simply O( t), while the numerical scheme is
said to be first order accurate.

As the numerical integration proceeds in time, the errors of each individual time step
are accumulated. It should be remarked here that the accumulative error for any t is
subject to a decay by virtue of the meaning of Equation (2.1) and for any realistic
time step t the accumulated error will tend to zero for t à . Another interesting
observation regarding the truncation error in the numerical integration of Equation
(2.1) is the nearly-linear decrease of the error when reducing the time step t from
2.0 days to 0.5 days, as shown in Table 2. 1. This behaviour points at a rapidly
decreasing influence of the higher-order derivatives in the truncation error, for this
application.

The much smaller truncation error in the implicit scheme of Equation (2.7) can also
conveniently be demonstrated by a Taylor's series expansion. This derivation
follows a more common introduction of Taylor's series in numerical schemes. After
the selection of the appropriate centre point of the scheme all values of the
dependent variables introduced at neighbouring grid points are expanded from that
centre point. For the implicit scheme, centred at point n+½, the Taylor's series
expansion gives

æ dh ( ½Dt ) d 2 h (½Dt ) d 3 h ö
2 3

(1 + ½aDt ) çç h + ½Dt + 2
+ 3
+ h.o.t . ÷
÷
=
è dt 2 dt 6 dt ø
æ dh ( -½Dt ) d 2 h ( -½Dt ) d 3 h ö
2 3

(1 - ½aDt ) çç h + ( -½Dt ) + 2
+ 3
+ h.o.t . ÷
÷
è dt 2 dt 6 dt ø

where all values of h and the derivatives with respect to time are taken at grid point
n+½. Expanding these expressions further leads to the equation

dh a d 2h 1 d 3h
= - a h - Dt 2 2 - Dt 2 3 + h.o.t. (2.14)
dt 8
144444dt 424 dt
2444444 3
-TE

which is the equation in reality solved by the implicit scheme.

WL | Delft Hydraulics 25
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Considering, again, that for this application all third- and higher-order derivatives
are small, the remaining part of the truncation error is indeed much smaller than that
of Equation (2.12) for any realistic choice of t. A realistic time step in this context
is a choice which relates t to the value of as discussed in § 2.10.

Figure 2.5 Oscillations and instabilities generated for large time steps: (a) exact solution; (b) stable
oscillatory solution for t=18 days; (c) unstable solution for t=22 days

2.8 Stability
Despite the truncation error in the computation presented in Table (2.1) most results
are quite acceptable for practical purposes. However, it is interesting to observe
what kind of results would have been produced if the time step had been taken much
larger. As an example, let us consider a time step of 25 days in an application of the
Euler scheme for the same equation and data as used in Table (2.1). For successive
time steps the sequence of results would be 10, -15, 22.5, -33.75, 50.63, -75.94 etc.,
leading to infinity or an exponent overflow message on a digital computer after a
sufficiently large number of time steps. In any hand computation the sequence of
operations would have been interrupted after one or a few time steps as the results
would appear to be unrealistic for any practical interpretation. A computation of this
kind is called an instability. In an unstable computation results will always exceed a
limit which has been set by the engineer as a realistic maximum or minimum value,
for which exceedance is not to be expected (Figure 2.5). A frequently used analysis
for the definition of stability criteria is based on the notion of amplification factors
between results at successive time steps. If the absolute value of the amplification
exceeds unity at each and every step in time, one has sufficient proof of the unstable
nature of the computation. The application of this analysis to the Euler scheme of
Equation (2.4) gives

h n +1
| A| = = 1 - aDt £ 1 (2.15)
hn

as a stability condition for the scheme. Since t is definite positive the stability
condition for the Euler scheme is derived as t 2.

WL | Delft Hydraulics 26
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Applying the same reasoning to the implicit scheme of Equation (2.7) leads to

1 - ½aDt
£ 1 (2.16)
1 + ½aDt

and to the conclusion that the implicit scheme is unconditionally stable. However,
the stability limit of t=20 days obtained for this application of the Euler scheme is
far beyond any time step that would be set as a maximum from the accuracy point of
view. Even for the implicit scheme the results obtained with this time step are very
inaccurate, as h drops to zero over the first time step and remains zero over all
successive time steps, whereas in the exact solution the value of h decreases
exponentially from the given initial value and only approaches the value of zero for
tà .

It may be included that the unconditional stability of the implicit scheme does not
have special advantages in this application to the solution of ordinary differential
equations. When moving to applications on partial differential equations, however,
the increased stability of the implicit schemes will turn out to be of such great
importance that currently, nearly all finite difference schemes are based upon
implicit formulations.

2.9 Consistency and convergence


The Taylor's series expansion of the numerical scheme visualizes the difference
between the differential equation as a continuous description of the physical system
and the finite difference equation as a discrete description on a set of grid points. As
t decreases, the difference between both equations reduces to the extent that they
become equivalent as tà0. In such case the difference equation is said to be
consistent With the differential equation. In the case of ordinary differential
equations this generally implies that the results also converge towards the exact
solution as tà0. An additional condition for such convergence is that the results
are computed under stable conditions.

2.10 The choice of a time step


The choice of a time step in the numerical integration of the differential equation is a
balance between the maximum tolerable error and the economy of the numerical
operations. In any model application, errors are introduced from the following
sources :

· accuracy of basic data;


· choice of additional parameter values;
· model schematization;
· choice of simulation data;
· numerical errors.

WL | Delft Hydraulics 27
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Of all these sources of errors the numerical error is easiest controlled and a general
requirement in model simulations is that the numerical error does not add to the
uncertainties in the results introduced by the other sources. To satisfy this condition,
the numerical error should be of an order of magnitude smaller than the overall
expected error. So, if the overall error is expected to be some centimetres in level,
the admissible numerical error should not exceed a few millimetres. Whereas it is
already difficult to estimate the magnitude of the overall error, it is even more
difficult to estimate the numerical error generated during one single time step and
even more so the accumulated effect over various steps.

As the truncation error contains higher-order derivatives, a first estimate of the time
step is based on an idea about the curvature of the solution function. Strongly curved
solution functions require smaller computational steps than solutions with more
gentle variations.

Figure 2.6 Process of diminishing errors by reducing the time step t

A better estimate of the time step is based on a sensitivity analysis. By taking


successively smaller grid steps, solutions are compared and if the differences appear
to be sufficiently small, similar computations can be made with that acceptable grid
step. It should be noted that the difference obtained by successively halving the
time step is less than half the total numerical error , as shown in Figure (2.6), where
this total numerical error at a given point in time is plotted as a function of the
number of computational time steps N over a given period of integration. This leads
to the conclusion that it is more efficient to refine coarse grids than refining further
already fine grids.

In a pragmatic approach one might also limit the allowable change in the function
derivative over a single time step. For the simple Equation (2.4), this condition is
equivalent to setting a maximum to the change in the value of h from one time step
to the other. Setting, for example, as a criterion that over a single time step a change
of 5% is allowed, this criterion leads to t 0.05, or t 0.5 days.

Even in this simple case, however, it remains difficult to estimate the accumulated
effect of this error over various time steps. In an attempt to do so and keeping in
mind the decaying nature of the error, the accumulated error E at time step n=N, is
approximated as

å ½ Dt a h n (1 - aDt )
N -n
E = 2
(2.17)
n =1

WL | Delft Hydraulics 28
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where in this notation N-n represents an exponent. In principle, it is possible to


follow the magnitude of E during the computational process, assuming that the
effect of third- and higher-order derivative contributions to the truncation error are
negligible. Such an approach, however, is not practical for more complex problems
and it is better to turn to sensitivity analysis for determining an acceptable time step.

2.11 Reservoir routing


Consider a reservoir with a level-dependent storage area A, an inflow Qi given as a
hydrograph and a level-dependent spillway outflow Q0, as shown in Figure (2.7).

Figure 2.7 Sketch of a reservoir with spillway flow

Reservoir volume balance and spillway flow are given by the following set of
equations

dh
A = Qi ( t ) - Qo (2.18)
dt

Q0 = 1.71 m L ( h - hcr )
1.5
(2.19)

where
A - level dependent surface area of the reservoir;
h - reservoir water level above a general reference (e.g. mean sea level
MSL);
hcr - level of the spillway crest;
L - length of the crest;
m - discharge coefficient.

In principle, Equation (2.19) may be substituted into Equation (2.18) to give one
single equation with h as the only dependent variable. However, in general it is
preferred to do such substitutions at the level of the numerical formulation after
linearization of the equation and/or the finite difference formulation.

In a simple Euler approach the equation reads as follows:

= 1.71 m L ( h n - hcr )
1.5
Q0 n (2.20)

WL | Delft Hydraulics 29
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Dt
h n +1 = hn +
An
{
Qi ( t n ) - Qo n } (2.21)

where n, again, has the meaning of a superscript indicating the grid point along the
time axis. To demonstrate the computational algorithm, a small example is worked
out with the following data:

t Qi(t) h A(h)
hinitial= 24.70 m (hours) (m3/s) (m) m2
hcr= 24.00 m 0 50 24 0.4*106
L = 30 m 1 150 25 0.8*106
t = 1 hour 2 360 26 1.0*106
m = 1.1 3 340 27 1.1*106

The results of the computation are given in Table (2.2).

Table 2.2 Results of the reservoir routing simulation with the Euler method

time grid hn Qon Qin An h


(hrs) point (m) (m3/s) (m3/s) (m2) (m)
0.0 0 24.700 33.05 50 6.80*105 0.090
1.0 1 24.790 39.62 150 7.16*105 0.555
2.0 2 25.345 88.02 360 8.69*105 1.127
3.0 3 26.472 219.32 340

With reference to Figure (2.8) it is readily seen that the time step of 1 hour is too
large, as the error introduced by using a reservoir surface area at time n t is
significant. Moreover, the use of the discharge at time n t contributes to the error in
the time derivative, although it does not affect the volume balance in a direct way.

Figure 2.8 Volume error at successive time steps in reservoir routine using the Euler scheme

A much improved formulation is based on the Newton-Raphson approach. Apart


from the better centring of the derivative of Equation (2.23) the variation in the
surface area is included implicitly. Although this variation is only introduced as a
linear function, it is a great improvement over including A as a constant over a
computational step, defined at time level n t.

WL | Delft Hydraulics 30
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

For the Newton-Raphson formulation the equations are rewritten in the form

dh 1
dt
=
A
( Qi ( t ) - Qo ) = F ( h, Qi , Qo ) (2.22)

and discretized, jointly with Equation (2.20), as

Dh
= F n + ½DF (2.23)
Dt
n
æ dQ ö
Qo + DQo
n
= Qo ( h ) + ç o ÷ Dh
n
(2.24)
è dh ø
where
¶F ¶F ¶F
DF = DQi + DQo + Dh
¶Qi ¶Qo ¶h
n (2.25)
=
1
A n (
DQi - DQo ) -
1
( An )
2 ( ) æ dA ö
Qi ( t ) - Qo n ç ÷ Dh
n

è dh ø
and
dQo
= 1.5*1.71* m L ( h - hcr )
0.5
(2.26)
dh

Special attention is drawn to the use of Q0(hn) in Equation (2.24) instead of


substituting the already known value Q0n. As shown in Figure (2-9) the value Q0n
was obtained from a linearized Q-h relation at time level (n-1) t (line a at point A).
The solution should proceed from point B along the line b during the next time step.
If in the right hand side of Equation (2.24) Qon had been used instead of Qo(hn), the
solution would proceed from point C along the dotted line c and the accumulation of
errors over various time steps would bring us further and further away from the
correct Q-h relation. Note that for consistency reasons this correction is not included
in the volume balance equation.

Figure 2.9 Linearization of the Q0-h relation

WL | Delft Hydraulics 31
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The coefficients in the two linear equations can be collected to the form

a11Dh + a12 DQo = b1


(2.27)
a21Dh + a22 DQo = b2
or
- -
Ax = b (2.28)
- -
with matrix A and vectors x and b given as

éa a12 ù - é Dh ù - éb ù
A = ê 11 ; x = ê ú; b = ê 1ú (2.29)
ë a21 a22 úû ë DQ0 û ëb2 û

where

a11 =
2 An 1
Dt A
( æ dA ö
+ n Qi ( t n ) - Qo n ç ÷
è dh ø
)
a12 = 1
n
æ dQ ö
a21 = ç 0÷ (2.30)
è dh ø
a22 = -1
b1 (
= 2 Qi ( t n ) + ½DQi - Qo n )
b2 = Qo n - Qo ( h n )

Elimination of Q leads to

b1 + b2
Dh = (2.31)
a11 + a21

DQ = b1 - a11Dh (2.32)

The computation over the same time steps as taken for the demonstration of the
Euler method gives results as shown in Table (2.3). The results, indeed, are more
realistic than those of Table 2.2. However, the differences between Q0n and Q0(hn)
indicate that it is advisable to use a smaller time step of integration.

WL | Delft Hydraulics 32
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Table 2.3 Results of the reservoir routing simulation with the implicit Newton-Raphson method

time (hrs) grid hn (m) Q0n (m3/s) Q0(hn) (m3/s) Qi(t) An*105 dA/dh dQ0/dh
point (m3/s) (m2) *105 (m) (m2/s)
0.0 0 24.700 33.05 33.05 50 6.80 4.00 70.82
1.0 1 24.992 53.73 55.75 150 7.97 4.00 84.31
2.0 2 25.688 114.45 123.78 360 9.38 2.00 109.98
3.0 3 26.379 199.71 207.02 340

time (hrs) grid a11 a21 b1 b2 h (m) Q0 (m3/s)


point
0.0 0 388 70.8 133.9 0.00 0.292 20.68
1.0 1 491 84.3 402.5 -2.03 0.696 60.72
2.0 2 573 110.0 481.1 -9.33 0.690 85.27
3.0 3

2.12 Routing of a decayable pollutant through a reservoir


Let us consider next a similar reservoir which is polluted by a decayable substance
with concentration c. Assuming that the reservoir volume can be schematized as
well mixed and introducing a first order decay, with reaction coefficient k, gives

d
( cV ) = ci Qi - c Qo - k cV (2.33)
dt

or
dc dh
V ( c ) + cA = ci Qi - c Qo - k cV (2.34)
dt dt

Substitution of Equation (2.18) into Equation (2.34) and division of all terms by the
reservoir volume V leads to

dc Qi
= ( ci - c ) - kc (2.35)
dt V

where the volume V, as a function of the reservoir level, follows from the integration
of the surface area A along h. Such integration is best based upon a simple
trapezium rule as the surface area, obtained from planimetering a topography from a
map, is never accurate enough to justify higher-accuracy integrations of the A-h
relation, such as the Simpson's rule or even integrations based on cubic spline
functions. Moreover, linear A-h relations are used in the Newton-Raphson method
and the use of the trapezium rule for the integration of the V-h relation is consistent
with this approach. Although in a quick analysis the dilution coefficient
Qi
D = (2.36)
V
is often set as a constant, giving for Equation (2.35) the exact solution

D
c = c0 e - ( k + D ) t - ci ( e - ( k + D )t - 1) (2.37)
k+D

WL | Delft Hydraulics 33
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

We will focus here on the general case where D is a function of time and the exact
solution of Equation (2.35) does not exist. The simplest numerical solution of
Equation (2.35) is, again, based on the Euler method, giving

æQ ö
= c n + Dt ç i ÷ ( ci - c ) - kc n Dt
n
c n +1 (2.38)
èV ø

Numerical solutions have the advantage over analytical solutions that all parameters
can easily be made a function of the dependent variables c and h and of the
independent variable t.

Examples of such further generalization of relations both in the water quantity and
quality part are:

· outflow partly given as a user demand, possibly affected by reservoir operation


levels;
· inclusion of the effects of precipitation to and evapotranspiration from the
reservoir surface in a longer term water balance simulation;
· stage dependent width of the spillway crest;
· controlled spillway crest level as a function of h or t;
· the use of m, calibrated as a function of the reservoir level;
· various sources of pollutant inflows into the reservoir;
· decay described as a function of the time-varying water temperature;
· inclusion of additional substances, such as dissolved oxygen with dependence on
wind-generated reaeration, photosynthesis, bottom sediment processes etc.

Again, in this example more accurate results are obtained for a given time step with
the improved Euler method and implicit methods, including the Newton-Raphson
formulation. As the water balance is not affected by the pollutant concentration, any
implicit technique leads to a system of linear equations with the unknown
concentration at the new time level decoupled from the unknown level and out
flowing discharge. It is also rather common to simulate the water quantity part first,
write results to a data base and subsequently retrieve the necessary information in a
separate water quantity computation. We will return to this approach when
discussing water quality studies in rivers, coastal areas and in reservoirs where the
assumption of a well mixed estuary is not correct.

2.13 Non-uniform steady flow in channels


As an example of computations which describe steady processes with variations in
the spatial x-direction let us consider the backwater computation in a uniform
channel. A typical channel cross-section is shown in Figure (2.10), including the
conveyance K plotted as a function of the water level, where
jj
1
K = åj =1 nj
Aj R j 2 / 3 (2.39)

WL | Delft Hydraulics 34
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

with

nj = Manning roughness coefficient for the sub section j;


Aj = sub area of the cross-section j;
Rj = local hydraulic radius of the sub area j;
jj = number of vertical slices used in the integration across the section.

Figure 2.10 Cross-section and cross-sectional parameters used in the non-uniform flow computation

Assuming the absence of lateral flow, the steady state De Saint Venant equations
reduce to the form

Q = Qcons tan t (2.40)


1 d æ Q ö dh
2
Q 2

ç ÷+ + I0 + 2 = 0 (2.41)
gA dx è A ø dx K

Further differentiation of the first term of Equation (2.41) gives, for constant Q,

æ Q 2 dA ö dh Q2
ç1 - 3 ÷ + I0 + 2 = 0 (2.42)
è gA dh ø dx K

or
dh 1 æ Q2 ö
= I
ç 0 + ÷ (2.43)
dx Fr 2 - 1 è K2 ø

with the dimensionless Froude number Fr defined as a function of the flow velocity
u, the cross-sectional area A and the storage width b, as

u2 Q 2 dA
Fr 2 = = (2.44)
A gA3 dh
g
bs

For a given discharge, the right hand side of Equation (2.43) is a function of the
water level above the channel bottom, giving

dh
= F ( h) (2.45)
dx

WL | Delft Hydraulics 35
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Table (2.4) gives an example of such function F(h), where it is readily seen at which
depth the equilibrium slope is reached. This depth, for which F(h) = 0, is called the
normal depth.

The solution can easily be found in a Newton-Raphson approach, where, following


the usual notation, a subscript i has been introduced as a grid point counter along the
x-axis. The implicit finite difference scheme reads as

Dhi æ dF ö
= Fi + ½ ç ÷ Dhi (2.46)
Dx è dh øi

or
Fi Dx
Dhi = (2.47)
æ dF ö
1- ½ ç ÷ Dx
è dh øi
For the cross-sectional data given in Figure (2.10), a discharge Q=3.40 m3/s, a bed
slope I0=-10-4, the uniform flow depth is found for dh/dx=0, or Q2=-K2I0, giving
K=340 m3/s and h=1.00 m from the conveyance function shown in Table (2.4). At
the downstream end of the channel the flow is controlled by a weir with a free
overflow given by Equation (2.19) with parameter values:

m = 1.1 (weir coefficient);


L = 11.0 m (weir crest length);
hcr = 2.00 m (weir crest level),

giving a water level above the bottom just upstream of the weir of h0=2.30 m. The
value serves as the downstream boundary condition for the non-uniform flow
computation. In the situation sketch of Figure (2.11) attention is drawn to the fact
that the numerical integration of Equation (2.43) proceeds in the negative x-
direction, requiring a negative value for the distance step x.
A suitable choice for this distance step follows from the consideration that in the
first integration step the change in h should be limited to a reasonably small fraction
of the difference between the downstream boundary water depth and the uniform
flow depth. Considering that the Newton-Raphson approach gives fairly accurate
results, a first step h = 0.10 m, covering approximately 10 % of the total difference,
should be acceptable. Linear interpolation in Table 2.3 gives a value K = 1498 m3/s
at the downstream boundary. For a friction slope Q2/K2 = 0.052 * 10-4, giving a
difference of 0.95 * 10-4 with the bed slope, a choice x = -1000 m satisfies the
criterion selected. Completing the table with the water level gradient function F of
Equation (2.43) for the given discharge (Table 2.4), the upstream water depths can
be computed by applying Equation (2.47). Results are shown in Table (2.5) over a
distance of 12 km upstream of the weir.

WL | Delft Hydraulics 36
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 2.11 Positive backwater upstream of a hydraulic structure

Table 2.4 Cross-sectional data given as a tabulated function of the depth, including discharge
dependent functions for Q = 3.4 m3/s (see also Figure 2.10)

water bs (m) A (m2) K (m3/s) Q2/K2 *10-4 Fr F *10-4 (dF/dh)


depth h *10-4
(m) (m-1)
1.0 10.0 9.00 340 1.000 0.126 0.000
1.512
1.5 11.0 14.25 688 0.244 0.066 0.756
0.312
2.0 12.0 20.00 1143 0.088 0.042 0.912
0.100
2.5 20.0 28.13 1734 0.038 0.032 0.962

Table 2.5 Non-uniform flow computation based on the Newton-Raphson implicit approach (a)
and compared with results of the Euler method (b)

(a) Newton- Raphson (b) Euler


x grid hi Fi (dF/dh)i h hi Fi h
(m) point i (m) *10-4 *10-4 (m) (m) *10-4 (m)
(m-1)
0 0 2.300 0.942 0.100 0.369 2.300 0.942 -0.094
-1000 1 2.206 0.933 -0.093
-2000 2 2.113 0.923 -0.092
-3000 3 2.021 0.914 -0.091
-4000 4 1.930 0.890 0.312 0.335 1.930 0.890 -0.089
-5000 5 1.841 0.862 -0.086
-6000 6 1.755 0.836 -0.084
-7000 7 1.671 0.809 -0.081
-8000 8 1.595 0.786 0.312 0.296 1.590 0.784 -0.078
-9000 9 1.512 0.760 -0.076
-10000 10 1.436 0.659 -0.066
-11000 11 1.370 0.559 -0.056
-12000 12 1.327 1.314

The results are also computed by using the Euler scheme which differs from the
Newton- Raphson scheme by the second term in the denominator of Equation (2.47).
Assuming that the Newton-Raphson results are rather accurate, for this distance step,
the Euler scheme produces results which might be considered acceptable for
practical use. Although the difference of 1.3 cm at x=-12000 m will still increase
further upstream, the cross-section schematization, the constant bed slope and the
choice of roughness coefficients will give rise to significantly larger uncertainties
about the correct water levels.

WL | Delft Hydraulics 37
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

For another observation we refer to Table (2.4). The number of levels for which the
cross- sectional parameters are given is fairly small and produce unacceptable errors
in the linear interpolation of these values. The errors introduced are most likely
larger than those caused by using a lower accuracy numerical integration scheme.
Although this might be acceptable in a quick analysis by hand computation, any
computer code would be programmed to set up the table with much smaller steps.
This table step size should not exceed substantially the step h in the numerical
integration. For our problem it would have been advisable to present the functions
bs, A, K, Fr and F at intervals of 10 cm.

The advantage of the Newton-Raphson scheme over the Euler scheme is the higher
accuracy obtained at the cost of the simple additional computation of the
denominator of Equation (2.47). For a comparison of its efficiency with that of the
normal Euler scheme let us recompute the solution with a four times larger distance
step, as shown in Table (2.6).

Table 2.6 Recomputation of the backwater curve with x=-4,000 m

(a) Newton- Raphson (b) Euler


x (m) grid hi Fi (dF/dh)i h hi Fi h
point i (m) *10-4 *10-4 (m) (m) *10-4 (m)
0 0 2.300 0.942 0.100 0.369 2.300 0.942 -0.377
-4,000 1 1.931 0.890 0.312 0.335 1.923 0.888 -0.355
-8,000 2 1.596 0.786 0.312 0.296 1.568 0.777 -0.311
-12,000 3 1.300 1.257

The Newton-Raphson approach still maintains its high accuracy during the first two
steps. During the third step the effect of the large table step in felt in the correction
through the term dF/dh. The difference of 1.7 cm with the computation taking x =-
1,000 m, would certainly have been much smaller if Table (2.4) had been set up with
smaller steps h. The result obtained with the Euler method for this case differs 7.0
cm from the most accurate result and is quite unacceptable.

The advantage of the Newton-Raphson scheme becomes clearer when negative


backwaters with much stronger water profile curvatures are computed. Such
computations start from downstream boundary conditions with depths below the
normal depth h=1.00m. There is no doubt that the Newton-Raphson scheme is far
more efficient in its application than the Euler scheme.

2.14 An inverse scheme for the backwater curve


computation
It has been demonstrated that once the table of the function F of Equation (2.45) has
been constructed for a uniform cross-section, the numerical operations for the non-
uniform flow computation are fairly simple. For uniform channels with a constant
bed slope we will now introduce a scheme which combines a high accuracy with an
extreme simplicity. The computation of depths with the Newton-Raphson method
introduced in the previous section is an asymptotic process, requiring continuous
computations with ever decreasing changes in the depth.

WL | Delft Hydraulics 38
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Of course, one could have decided to increase the distance step as the depth comes
closer and closer to the normal depth.

However, returning to Equation (2.45), one could also introduce a numerical scheme
taking equal steps h in the dependent variable h, a so-called inverse scheme. The
advantage of this approach is that the right hand side of the equation, being only a
function of h, can be centred without requiring the iteration of the improved Euler
scheme. The simple inverse scheme looks as follows:

Dh
Dxi = (2.48)
Fi +1/ 2

and results are shown in Table (2.7) for steps h = 0.10 m. These results are based
on the same interval for presenting topographic data as given in Table (2.4).
Interpolation at x = - 8,000 m in Table (2.7) gives a depth of 1.596 m, which only
differs 0.1 cm from the result obtained with the Newton-Raphson scheme. At x = -
12,000 m, however, the difference is 0.5 cm, which difference must mainly be
attributed to the Newton-Raphson integration based on too large table intervals (grid
point 9-10).
The normal depth h=1.00 m is found at x = -33 km, whereas theoretically this should
be at x = - . However, this deviation is of no practical importance at all. Although
the inverse scheme is very handy for uniform channel computations, it looses its
advantage or even does not work at all in applications where the channel parameters
vary along the channel axis.

Table 2.7 Results of the inverse scheme for h = 0.10 m

i xi (m) hi (m) Fi+½ * 10-4 xi (m)


0 0 2.300
2.25 0.9370 -1,067
1 -1,067 2.200
2.15 0.9270 -1,079
2 -2,146 2.100
2.05 0.9170 -1,091
3 -3,237 2.000
1.95 0.8964 -1,116
4 -4,353 1.900
1.85 0.8652 -1,156
5 -5,509 1.800
1.75 0.8340 -1,199
6 -6,708 1.700
1.65 0.8028 -1,246
7 -7,954 1.600
1.55 0.7716 -1,296
8 -9,250 1.500
1.45 0.6804 -1,470
9 -10,720 1.400
1.35 0.5292 -1,890
10 -12,610 1.300
1.25 0.3780 -2,646
11 -15,256 1.200
1.15 0.2268 -4,409

WL | Delft Hydraulics 39
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

12 -19,665 1.100
1.05 0.0756 -13,228
13 -32,893 1.000

2.15 Non-uniform flow in non-uniform channels


Let us now consider the somewhat more complex case of steady flow in non-
uniform channels. For such channels the cross-sectional area is a function of both h
and x, giving the variation in the x-t space

¶A ¶A
dA = dx + dh (2.49)
¶x ¶h
or
dA ¶A ¶A dh
= + (2.50)
dx ¶x ¶h dx

Equation (2.43) should now read as

dh 1 æ Q 2 Q 2 ¶A ö
= ç I 0 + - ÷ (2.51)
dx Fr 2 - 1 è K 2 gA3 ¶x ø

For the solution of this equation it is most practical to process the function F (x, h)
as a tabulated function at equal depth steps h, starting from a reference point near
the channel bottom. It is advisable to choose a h that has the same value at all
cross-sections given along the river. For each river section, between two successive
cross-section points, a set of parallel lines is found along which F (x, h) reflects the
change in F along x, while keeping h constant. This approach will require at each
cross-section the processing of one table reflecting the values dA/dx applicable to
the left hand channel section and another table reflecting their values applicable to
the right hand section.

The Newton-Raphson integration now proceeds with the scheme

Dhi æ ¶F ö æ ¶F ö
= Fi + ½ ç ÷ Dx + ½ ç ÷ Dhi (2.52)
Dx è ¶x øi è ¶h øi
or
æ ¶F ö
Fi Dx + ½ ç ÷ Dx
2

è ¶x øi
Dhi = (2.53)
æ ¶F ö
1- ½ ç ÷ Dx
è ¶h øi

in an integration performed over an integer number of distance steps along each


successive channel section marked by its end cross-sections and its local reference
slope. Again here, the procedure is most efficient if only one Newton-Raphson
iteration is made per time step and care is taken that the distance steps are not too
large.

WL | Delft Hydraulics 40
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Some final remarks:

1. Both methods applied to uniform and to non-uniform channels can easily be


extended to include the effects of lateral flows.
2. Both methods apply to fully irregular cross-section geometry and allow for a
variation of roughness coefficients across the channel.
3. The methods are applicable for any formulation of the friction laws entered into
the integration of the channel conveyance K.
4. The efficient inverse scheme of Equation (2.48) can, in general, not be applied to
non-uniform channel flow computations.
5. Both methods can be applied to sub critical flow (Fr<1) and to supercritical flow
(Fr>1), if the appropriate direction of integration (algorithmic structure) is
followed.

2.16 What have we learnt?


In this Chapter we introduced various numerical methods for the solution of (sets of)
ordinary differential equations. Many of these problems can simply be programmed
in Excel. This is very straightforward in cases where equation parameters are
constants. Cases with non-linear parameters can still be programmed easily by using
macros for the function descriptions.

When programming solutions for incidental application it is recommended to use the


simple Euler method and to make sure that a sufficiently small grid step is used.
Higher accuracy methods only justify the additional effort of programming and
testing of the code when repetitive use of this code is made and a single computation
is not finalized instantaneously. With the current computer speed this is no longer
the case for the relatively simple problems programmed in Excel.

The higher accuracy integration methods have been introduced primarily, as the
principles demonstrated are commonly applied in the numerical methods for solving
partial differential equations. Applications of iterative methods as applied in the
improved Euler method, implicit schemes and Newton Raphson principles, will be
discussed in Chapters 4, 5 and 6.

2.17 Questions and small assignments


1. Extend the computation of all cases shown in Table (2.1) up to T = 6.0 days.

2. Explain in your own words why the improved Euler method is more attractive in
use than the standard Euler method.

3. Determine the type of numerical scheme used in the well-known reservoir


routing equation

DS 1
= ( I1 + I 2 - O1 - O2 )
Dt 2

WL | Delft Hydraulics 41
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where
S = reservoir storage ;
I = reservoir inflow ;
O = reservoir outflow,

with subscripts 1 and 2 referring to time levels n t and (n+l) t respectively.

4. Explain why the resulting functions in Table (2.1), both obtained with the Euler
method and the implicit scheme, is below the exact solution. Also explain why
the resulting function obtained with the improved Euler method lies above the
exact solution.

5. Find with a computation carrying more decimal positions than shown in Table
(2.1) the ratio between the errors obtained with the Euler scheme and the implicit
scheme for a time step of 1 day. Discuss this ratio in the light of the truncation
errors for both schemes.

6. Verify Equation (2.14).

7. Show with a more precise computation of the results of the implicit scheme
computation in Table (2.1) that at T = 4 days the error obtained with t = 2 days
is four time larger than the error obtained with t = 1 day. Explain this
difference from the truncation error of the scheme as shown in Equation (2.14).

8. Give the fourth order derivative term of the truncation error of Equation (2.14).

9. Define an Euler scheme for Equation (2.1) based on a numerical integration in


the negative time direction and show that the scheme is unconditionally unstable.

10. Extend the computation shown in Table (2.2) with one additional time step.

11. Extend the computation shown in Table (2.3) with one additional distance step.

12. Extend the computation shown in Table (2.5) with one additional distance step.

13. Explain why the depths calculated with the Euler scheme in Table (2.5) are
smaller than those obtained with the Newton-Raphson scheme.

14. Explain why in Table (2.6) the depth obtained in a computation with a distance
step of 4 km is smaller than the depth obtained with a 1 km step.

15. Recompute the solution of Table (2.7) with h = 0.05 m over the range 2.1 h
2.3 m. Comment on the results.

16. Explain what sort of problems may result from the application of Equation (2.48)
for the computation of non-uniform channel flow.

WL | Delft Hydraulics 42
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

3 Basic Unsteady Channel Flow Equations

3.1 Introduction
Unsteady channel flow will be first derived for a simple uniform channel of unit
width, a horizontal bottom with friction along the walls and the bottom neglected.
Referring to Figure (3.1), the water depth is symbolized by h and the average flow
velocity by u. These variables u and h are usually called the dependent variables,
defined as functions of the independent variables x (space coordinate) and t (time
coordinate). The definition of two dependent variables requires the formulation of
two equations for their solution. In channel flow these equations are usually based
upon volume and momentum conservation. The volume conservation approach
assumes incompressible flow and constant water density. This is a quite realistic
assumption for free surface fresh water channel systems. The equation following
from this volume conservation principle is called the continuity equation.

Figure 3.1 Control volume along a channel axis

3.2 Continuity equation


The derivation of the equation is based on the concept that over a given time the
difference between inflow to and outflow from a control volume along the channel
balances the change in the storage in this control volume over this same time.
Referring, again, to Figure (3.1), it is readily seen that the inflow per unit time into
the control volume, or the flux of water volume through the upstream cross-section,
is given by

inflow = uh

Assuming that both u and h are continuous functions of x, the outflow at the
downstream section of the control volume can then be defined as

outflow = uh + (uh) dx
¶x

WL | Delft Hydraulics 43
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The storage in the control section per unit time is expressed in terms of the change in
water level as

storage = (h dx)
¶t

based on the concept that for continuous water level changes with initial depth h0,
the depth h1 after a step in time dt equals

¶h
h1 = h0 + dt
¶t
Balancing terms gives the continuity equation in the form

¶h ¶
+ (uh) = 0 (3.1)
¶t ¶x

Differentiating out the second term leads to another, often used, form

¶h ¶h ¶u
+u +h =0 (3.2)
¶t ¶x ¶x

3.3 Momentum equation


It should be recalled that momentum is defined as the product of velocity and mass.
The total momentum M of the fluid contained in our control volume of unit width,
therefore, is defined as
M = u r h dx

where M and u are both vectors acting along the channel axis. Over a certain step dt
in time, the change of M is balanced by the net amount of momentum carried by the
fluid velocity through the boundaries of the control volume and the impulse
generated by the forces acting upon it. Both quantities are to be taken with their
component acting in the x-direction. Following the same principles as used in the
derivation of the continuity equation, the net amount of momentum carried with the
flow velocity is defined by


M a = u r uh - [u r uh + (u r uh) dx ]
¶x

where the first term refers to the inflow of momentum and the term within braces to
the outflow of momentum per unit time.

We will furthermore recall Newton's second law

F dt = d(mu)

WL | Delft Hydraulics 44
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where F is a force acting on the fluid with mass m and velocity u. The product of the
force and the time period over which it is acting is called an impulse. Newton's
second law states that this impulse generates a change in momentum.

Figure 3.2 Forces acting on a horizontal uniform channel section

Under the simplifications introduced above, the only forces acting on the control
volume are the hydrostatic forces at the upstream and downstream ends of the
control section. Integrating the hydrostatic pressures over the verticals (Figure 3.2)
gives the net force

dF = ½r g h 2 - [ ½r g h 2 + ( ½r g h 2 ) dx ]
¶x
Balancing all terms over a unit time step leads, after division of all terms by the
constant dx, to the equation

¶ ¶ ¶
( r uh ) + ( ru 2 h ) + (½r g h 2 ) = 0
¶t ¶x ¶x

Dividing, furthermore, all terms by the constant , leads to

¶ ¶ ¶h
(uh) + ( u 2 h ) + gh = 0 (3.3)
¶t ¶x ¶x

Differentiating out the product terms in the derivatives, substituting the continuity
equation and dividing all the remaining terms by h, leads to

¶u ¶u ¶h
+u + g =0 (3.4)
¶t ¶x ¶x

3.4 Transformation to the characteristic form


Although the form in which the set of continuity and momentum Equations (3.1) and
(3.3) is shown is very convenient for the formulation of solutions on the basis of
finite differences, it is useful to introduce a transformation which enables us to
understand better the meaning of these equations. For this purpose we will introduce
the auxiliary variable

WL | Delft Hydraulics 45
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

c = gh (3.5)

Multiplication of Equation (3.1), substitution of Equation (3.3) it and division of all


terms by c gives

¶c ¶c ¶u
2 + 2u + c = 0 (3.6)
¶t ¶x ¶x
Similarly, the substitution of Equation (3.4)into Equation (3.3)s

¶u ¶u ¶c
+u + 2c = 0 (3.7)
¶t ¶x ¶x

The set of Equations (3.5), (3.6) represents the characteristic form of the flow
equations. The special significance of this form will become clear when respectively
Equation (3.5) is added to and subtracted from Equation (3.6), to give the set of
equations in the form

¶ ¶
(u + 2c) + (u + c) (u + 2c) = 0 (3.8)
¶t ¶x

and
¶ ¶
(u - 2c) + (u - c) (u - 2c) = 0 (3.9)
¶t ¶x

Before drawing any conclusions from these relations, we will introduce the general
mathematical formulation expressing the change of the value of a function f when
moving from a given point P to a neighbouring point Q.

Figure 3.3 Relation between function values at neighbouring points

As shown in Figure (3.3), this change results from the partial derivatives of the
function in t-direction and x-direction respectively as

¶f ¶f
df = dt + dx (3.10)
¶t ¶x

or, after division of both sides by dt

WL | Delft Hydraulics 46
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

df ¶f dx ¶f
= + (3.11)
dt ¶t dt ¶x

Taking for f the function u+2c, readily shows by comparison of Equations (3.7) and
(3.10) that along the line with direction

dx
= u+c (3.12)
dt
the expression
d
(u + 2c) = 0 (3.13)
dt

holds, while the comparison of Equations (3.8) and (3.10) shows that along the line
with direction

dx
= u-c (3.14)
dt

the expression

d
(u - 2c) = 0 (3.15)
dt

holds (Figure 3.4).

Figure 3.4 Characteristic directions and their Riemannn invariants

The line with the direction expressed by Equation (3.11) is called the c+
characteristic, whereas the line with the direction expressed by Equation (3.13) is
called the c- characteristic.

The integration of Equation (3.12) gives

u + 2c = J + = constant (3.16)

whereas the integration of Equation (3.14) yields the expression

WL | Delft Hydraulics 47
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

u - 2c = J - = constant (3.17)

In this formulation J+ and J- are the so-called Riemann invariants. For the simplified
channel flow equations these invariants have constant values along the so-called c+
and c- characteristics respectively. When introducing additional terms in the
equation, such as lateral flow, bottom slope and bottom friction, these Riemann
invariants will contain correction terms and will no longer be "invariant".

3.5 The significance of the characteristics


The conditions derived along the characteristics enable us to compute solutions from
known conditions at an earlier point in time. Referring to Figure (3.5), it is seen that
the state of the fluid at point P can be computed from given states at points A and B.
Given, for example, that at point A, u=1.0 m/s and h=5 m, while at point B, u=1.2
m/s and h=4.8 m, the Riemann invariants are computed as J+ = u+2Ögh = 15 m/s at
the points A and P, and as J- = u-2Ögh = -12.52 m/s at the points B and P. Solving
these two equations representing the Riemann invariants at point P gives u=1.24 m/s
and h=4.83 m.

Figure 3.5 Characterisitics leaving from points A and B, defining the solution of the flow state at a
point P

The significance of the characteristics, however, goes far beyond the possibility to
make such computations. The fact that along the characteristics a special condition is
valid, implies that they pass on some information on the state of the fluid. In other
words, the characteristics are to be seen as lines along which information on the state
of the fluid propagates. The uniqueness of the equivalence between Equation (3.10)
on the one hand and Equations (3.7) and (3.8) on the other hand, implies that these
lines along which information propagates are unique lines and that information only
travels along just these characteristics.

The auxiliary variable c introduced in Equation (3.4) appears to get a special


meaning as part of the direction of the characteristics. For the special case where
u=0, both characteristic directions have the same magnitude, though their signs are
opposite (Figure 3.6a). As c is a function of the water depth, this means that the
effect of any disturbance or control at a point along the river, propagates in both
channel directions with the same speed, which is a function of the square root of the
water depth. The variable c gets here the meaning of a celerity with which
disturbances propagate in stagnant water.

WL | Delft Hydraulics 48
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 3.6 Characteristic directions for: (a) and (b) subcritical flow; (c) critical flow and (d)
supercritical flow

For velocities u>0, the characteristic directions no longer have the same magnitude
and, at a particular point, information on the fluid state travels faster in the
downstream then in the upstream direction (Figure 3.6b). For increasing water
velocities this leads to the particular case where the celerity c equals the water
velocity u (Figure 3.6c) and at this point information is no longer able to travel in
the upstream direction. Introducing the Froude number, defined as

|u|
Fr = (3.18)
gh
it is readily seen that for the situation of Figure (3.6c), where u=c, the Froude
number equals unity. This special case is called critical flow, whereas the cases of
Figures (3.6a, 3.6b), where the Froude numbers are less than unity, represent sub
critical flow. For water velocities exceeding the celerity c, the Froude numbers
exceed unity and the flow is called super critical (Figure 3.6d).

From Figure (3.6) it is readily seen that for the case of sub critical flow, the state of
the fluid at any point of space and at any point in time is controlled by both the
upstream and the downstream conditions. In the case of super critical flow, however,
the state of the fluid is only controlled by the upstream conditions. So, if at any point
A at a channel, we have supercritical flow, downstream control of the conditions at
A is impossible, as long as the super critical state at A is maintained. It should be
noted, however, that this supercritical state may change into a subcritical state when
downstream control, by a weir for example, generates a region of subcritical flow
upstream of the control structure to the extent that this subcritical flow region
reaches point A.

From Figure (3.6) it may also be concluded that it takes a certain time before the
effect of the opening of the gate at the upstream end of an irrigation channel reaches
the point where the water is required. Taking, for example, a channel with stagnant
water at 1 meter depth, giving a celerity of 3.13 m/s, it lasts nearly 1 hour before
additional water reaches a demand point after the further opening of a gate 10 km
upstream.

WL | Delft Hydraulics 49
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 3.7 Regions of determinacy of the state at a point P for various flow conditions

To make these principles even more clear, Figure (3.7) shows various cases of
subcritical and supercritical flow, where the shaded areas show the region of
determinacy of the state at point P. Any disturbance or control within the shaded
region will have some effect on the water velocity and depth at point P, while any
disturbance or control outside the shaded region will not be able to affect conditions
at point P.
The propagation of information along the characteristics will also determine the
initial- and boundary condition requirements for the solution inside a computational
domain, bounded in the x-t plane by the lines t=0 and t=T and the lines x=0 and
x=L. Referring first to Figure (3.8a), point A at time t=0 receives two characteristics
coming from outside the computational domain. If point A would have been located
inside the computational domain, the solution would have followed from the
Riemann invariants along the two characteristics passing through it, much similar to
the situation sketched in Figure (3.5). With A located at the boundary of the
computational domain, it is can be concluded that the information on the Riemann
invariants has to be replaced by initial conditions, as we have no information on the
state of the fluid outside our computational domain and the values of the Riemann
invariants associated with it.

Figure 3.8 Relation between characterisitic directions and the boundary- and initial data
requirements

Similarly, at point B in Figure (3.8a), one characteristic arrives from inside the
computational domain and is supposed to carry information from another point
within the computational domain. The other characteristic arrives from outside the
domain, where the computation is not made and the information, which would
normally be passed on via the Riemann invariant, has to be replaced by a boundary
condition. Such boundary condition is usually provided in the form of a given
velocity, water level, discharge or stage-discharge relationship.

WL | Delft Hydraulics 50
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The same applies to point C, where one characteristic carries information from
inside the computational domain, while the other characteristic comes from outside
and requires a boundary condition to replace the information which would otherwise
be provided by its Riemann invariant.

Figure (3.8b) shows the case of supercritical flow with positive u. The problem is
completely upstream controlled and two upstream boundary conditions are needed
for a solution. We define this need as a two-point boundary condition against a one-
point boundary condition for the case of subcritical flow, given in Figure (3.8a).
Initial data requirements are the same as for the subcritical flow case and two-point
initial data are required for all types of flow. Usually, values for u and h are to be
given.

Figure (3.8c), finally, shows the case of a supercritical flow with velocity u in the
negative x-direction. For the same reasons as pointed out earlier, this system
requires two-point initial data and two-point boundary data at the upstream end,
which in this case is located at the point x=L.

From the discussion above the general conclusion can be derived:

one initial- or boundary condition is required for every characteristic


entering the computational domain,

where entering is defined by following the characteristic forward in time.

3.6 The method of characteristics


In § 3.3 we have introduced the Riemann invariants and the computation of the state
variables u and h with these Riemann invariants. Where necessary, these Riemann
invariants are complemented with the initial- and boundary conditions. We may now
proceed to the description of a complete solution algorithm, defined as the method
of characteristics.

This method is based on the construction of a network of points where the state
variables are computed. As shown in Figure (3.5), from a given set of two points on
the network a new point may be constructed by drawing a characteristic of one
family from one point, which intersects with a characteristic of the other family
drawn from the other point. In general, these directions are not known initially, as
they depend on the solution of the state variables at the intersecting point. However,
the solution of these state variables can be found through the Riemann invariants,
even if the exact location of the point is not known yet. The nearly exact direction of
these characteristics can then be drawn based on the average characteristic direction
at the connecting points.

The procedure is best explained by considering a practical example. Figure (3.9)


shows the application of the method on the problem of closing a gate in an irrigation
canal. Initially, the water flows at steady state with a velocity of 1.0 m/s at a depth of
1.60 meter. The gate is closed over a period of 100 seconds, controlled to give a
linearly decreasing water velocity immediately downstream of the gate.

WL | Delft Hydraulics 51
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

This assumption serves our exercise, as most gate control operations are based upon
water level sensoring. For a further simplification in this example it is also assumed
that the downstream channel is very long and that the wave, generated by the gate
closure, is not reflected at a downstream side.

Figure 3.9 Hodograph and physical planes for a graphical solution of the computations along
characteristics

For ease of computation the acceleration due to gravity is taken as g=10 m2/s. For
the initial steady state, the following data can be derived:

c+ = u+ Ögh = 5.00 m/s


c- = u- Ögh = -3.00 m/s
J+ = u+2Ögh = 9.00 m/s
J- = u-2Ögh = -7.00 m/s

The disturbance generated by the gate operation will travel in the downstream
direction with a celerity c+=5.00 m/s. This implies, for example, that at a point L
meter downstream of the gate, the flow remains undisturbed over a period of L/c+
seconds.

WL | Delft Hydraulics 52
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

As a result, one may expect a steady state region downstream of the gate, shown as
region A in Figure (3.9). This is the region enclosed by the lines (t=0; x 0) and
(x=5.00 t; x 0).

The solution will first be defined at points along the t-axis, immediately downstream
of the gate.
As a first step a set of points will be selected and numbered from 1 onward at
intervals of 25 seconds. Subscripts u and d will be used for the locations of these
points upstream and downstream of the gate respectively.

Let us first consider the solution of the problem downstream of the gate. The
solution at point 1d of Figure (3.9) is found by applying the boundary condition
u=0.75 m/s at t=25 s (Figure 4.9b), and the condition that the Riemann invariant J-=-
7.00 m/s at point 1d, as it keeps its constant value with which it arrives along the
negative characteristic from the steady state region A. The solution of this set of two
equations gives h=1.50 m.

Similarly, the solution at points 2d,3d and 4d is found by applying the velocity
boundary condition combined with the Riemann invariant carried along the negative
characteristic arriving from the steady state region A as:

point 2d: u=0.50 m/s & J-=-7.00 m/s --> h=1.41 m;


point 3d: u=0.25 m/s & J-=-7.00 m/s --> h=1.31 m;
point 4d: u=0.00 m/s & J-=-7.00 m/s --> h=1.22 m.

A summary of these data is shown in Table 3.1.

Table 3.1 Data at various characteristic net points of the physical plane

Point u h q gh c+ c- 2 gh J+ J-
(m/s) (m) (m2/s) (m/s) (m/s) (m/s) (m/s)
(m/s) (m/s)
0 1.0 1.60 1.600 4.00 5.00 -3.00 8.00 9.00 -7.00
1d 0.75 1.50 1.125 3.88 4.63 -3.13 7.75 8.50 -7.00
2d 0.50 1.41 0.705 3.75 4.25 -3.25 7.50 8.00 -7.00
3d 0.25 1.31 0.328 3.62 3.88 -3.37 7.25 7.50 -7.00
4d 0.00 1.22 0.000 3.50 3.50 -3.50 7.00 7.00 -7.00
1u 0.64 1.76 1.125 4.20 4.84 -3.56 8.40 9.00 -7.76
2u 0.38 1.86 0.705 4.31 4.69 -3.93 8.62 9.00 -8.24
3u 0.17 1.93 0.328 4.39 4.56 -4.22 8.79 9.00 -8.62
4u 0.00 2.03 0.000 4.50 4.50 -4.50 9.00 9.00 -9.00

Let us now turn to points x 0 away from the boundary and draw the positive
characteristic leaving from point 1d. Conditions at all points along this line are given
by the same Riemann invariants: J+ = u+2Ögh = 9.00 m/s and J- = u-2Ögh = -7.00
m/s. Consequently, the same solution is found at all points along this line, giving
u=0.25 m/s and h=1.50 m. This characteristic, then, is a straight line defined by a
celerity c+ = u1+Ögh1 = 4.63 m/s.

WL | Delft Hydraulics 53
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Similarly, the positive characteristics leaving from points 2,3 and 4 must be straight
lines with celerities of 4.25, 3.88 and 3.50 m/s respectively.
The region between the positive characteristics leaving from points 1d and 4d is
defined as a simple wave region, characterised by straight (positive) characteristics
of one family and curved characteristics of the other family. However, it is not
necessary here to draw the curved c- characteristics, as the visualisation of these
lines does not add any essential information on the state of the flow in such a simple
wave region.

For the same reason, only the characteristics bordering a steady state region are
drawn and none of the characteristics inside the steady state region.

Having defined a simple wave region one might define a complex wave region as
one where both families of characteristics are curved lines, indicating that velocities
and depths vary from point to point in both directions. In such a complex wave
region a network of both families of characteristics is drawn. This situation would
occur if the simple wave generated from points 1d to 4d would be reflected against a
downstream boundary as could easily be verified by introducing, for example, a
constant water level condition at a point, say, 400 m downstream of the gate. As will
be explained later, however, we do not want to carry the presentation of the method
of characteristics too far and refer the interested reader to standard text books on the
topic (e.g. Abbott, 1979).

However, as an example of a curved characteristic and for demonstration purpose


only, consider the negative characteristic between point 1d and the steady state
region A shown in Figure (3.9). The direction of this characteristic is varying from
point to point, as follows from the varying J+-values, while moving along this
characteristic from point 1d to region A. As a reasonable approximation the average
direction is taken as

c - 1d + c - A = -3.13 - 3.00 = - 3.07 m/s


c - 1d ® A =
2 2
and the approximate characteristic is drawn backward in time, leaving from point 1
until its arrival at the steady state region.

From this procedure it may be concluded that, although the method of characteristics
is generally considered a quite accurate solution of the simplified partial differential
equations describing channel flow, it does not represent an exact solution of these
differential equations. The approximation is in the location of the points of the
network. The solution of the simplified equations at the network points, however, is
exact. The more generally applied finite difference methods, to the contrary, have
exact locations of the network- or grid points, while the solution of the equations at
these points, as a rule, are approximations of the true solutions.

WL | Delft Hydraulics 54
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

3.7 More complex boundary conditions


Let us now consider the upstream part of the channel, where the water velocity is
reduced and the flux of momentum from upstream is likely to lead to a piling up of
the water against the control structure. The effect of any control will be travelling in
upstream direction with the celerity c- = -3.00 m/s.
As also in this case some time will elapse before the disturbance reaches an
upstream point along the canal, we will also find here a steady state region shown as
D in Figure (3.9).
The solution upstream of the gate requires gate boundary conditions. These are not
the same as given for the downstream part, as the velocities given are related to
cross-sections and water levels just downstream of the gate. The solution at this
downstream section, however, enables the combination of velocity and depth to give
the specific discharges, which are the same just upstream and downstream of the
structure.
The solution of the state at point 1u then follows from the computed specific
discharge q=udhd=1.125 m2/s and the positive Riemann invariant arriving from
steady state region D, J+ = u+2Ögh = 9.00 m/s. Substitution of one equation into the
other leads to a third degree polynomial in h, which requires a Newton-Raphson
iteration algorithm for their solution.

For non-computerized solutions it is therefore easier to turn to a graphical method.


Graphically, the solution of any set of two equations is found at the intersection of
the lines representing these equations in an appropriate coordinate system. For the
given problem, a suitable choice of the coordinates is a plot of values 2Ögh linearly
along the vertical axis and values u along the horizontal axis. The plane formed by
these two variables is called the hodograph plane. The choice of the vertical scale
leads to the representation of the Riemann invariants as lines under 45o (Figure
3.9c). By plotting the 2Ögh-axis vertically down, for subcritical flow the sign of the
directions coincides with that of the characteristic directions in the physical plane.
This convention is quite convenient in its practical use.

Table 3.2 Data for drawing q-relations in the hodograph plane

q u h gh 2 gh
m2/s m/s m
m/s m/s
1.125 .60 1.87 4.32 8.65
.70 1.61 4.01 8.02
0.705 .35 2.01 4.48 8.97
.40 1.76 4.20 8.39
0.328 .18 1.82 4.27 8.53
.16 2.05 4.53 9.06

WL | Delft Hydraulics 55
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure (3.10) also shows various boundary conditions, which may all easily be
presented in graphical form, such as u given (Figure 4.10c), h given (Figure 3.10d),
a given specific discharge q (Figure 3.10e) and a critical flow condition (Figure
3.10f). This last relation is a special form of a rating curve, or Q-h relation, which,
for the simplified channel flow equations, would normally be presented as a u-h
relation. The plot of a given specific discharge, as applied at point 2u of Figure
(3.9c), is easily made by selecting a number of u-values and calculating the
associated h- and 2Ögh-values. It should be realised that q should always be plotted
with the appropriate sign.
Also, the line is curved and requires a certain number of computed points in the
expected neighbourhood of intersection for a sufficient accuracy.

Figure 3.10 Various boundary conditions represented graphically in the hodograph plane

We return again to Figure (3.9c), which shows the hodograph plane belonging to the
physical plane of Figure (3.9a). The boundary conditions q=1.125, 0.705 and 0.328
m2/s at points 1, 2 and 3 respectively, are presented by three different lines in the
hodograph plane, whereas the discharge q=0 is represented by the line u=0. The
solutions at points 1u,2u,3u and 4u are shown in the hodograph plane as the
intersections of the lines representing the specific discharge at those points and the
condition J+=9.00 m/s. The characteristics leaving the t-axis over the period of the
closure, again, define a simple wave region, with characteristics now converging
towards each other. Further upstream this will lead to the intersection of these c-
characteristics. Such intersection of characteristics of the same family leads to two-
valued solutions for both u and h at such intersection, or, in other words, to the
formation of a hydraulic jump. It should be realised that the intersection of
characteristics is found here on the basis of the equations which do not contain bed
slope and bed friction. Where these terms are playing a significant role the
computations become more complex and are the subject of another chapter.

WL | Delft Hydraulics 56
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Although not frequent in nature or in man-made systems, such travelling hydraulic


jumps are known phenomena, for example, as tidal bores in estuaries or as undular
hydraulic jumps in tail race channels of peak power generating hydropower systems.

3.8 The formation of positive and negative hydraulic


jumps
The discussion on the formation of hydraulic jumps is simplest presented on the
basis of an extreme operation mode of the gate. Figure (3.11) shows the case of an
instantaneous closure of otherwise the same problem as discussed in Figure (3.9).

Considering first the upstream side of the gate, the positive Riemann invariant
carried from the steady state region D has to be combined with the condition u=0 at
x=0, for t>0. As shown in the hodograph plane the solution of these two conditions
is u=0 m/s; h=2.03 m. The negative characteristic celerity is c-=-4.50 m/s, which is
greater than c-=-3.00 m/s found over the steady state region D.
This difference in celerities leads to the immediate intersection of the negative
characteristic directions just upstream of the gate at t=0 and to the formation of a
positive hydraulic jump.

WL | Delft Hydraulics 57
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 3.11 Method of characteristics for the case of a sudden gate closure

We have assumed that despite this intersection of characteristics, the positive


Riemann invariant remains constant across the jump.
We will, furthermore, assume that the celerity of the jump is approximated by the
average of the two celerities of the intersecting characteristics forming this jump.
Both hypotheses have been investigated and appear to be approximately true for
small jumps (Abbott 1979). Based on these assumptions the celerity of the jump is
computed at c-=-3.75 m/s.

The line representing this propagation in the physical plane separates the steady state
region D from another steady state region C bordered by the line x=0, for t>0 and
the path of the hydraulic jump.

WL | Delft Hydraulics 58
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

A similar construction at the downstream side of the gate gives a steady state region
B, where h=1.22 m. Steady state regions A and B give characteristic celerities
c+=5.00 m/s and c+=3.50 m/s respectively, which show diverging directions
originating from point (x=0, t=0). Only at the very origin, a discontinuity is present,
which rapidly spreads along x while travelling downstream. We call such special
form of a discontinuity a negative hydraulic jump and the subsequent simple wave
region a centred simple wave. In this simple wave region the characteristics drawn
can be selected freely and Figure (3.11) shows the choice of some characteristics
representing steps of 10 cm in h.

The results of the computation along x, at T=125 s are shown in Figure (3.11c). The
discontinuity formed at the positive jump and the spreading negative hydraulic jump
in the downstream part are clearly recognised.

3.9 The limited practical importance of the method of


characteristics
The method of characteristics has the principal advantage of visualising the way in
which flow disturbances or the effects of flow control travel through the hydraulic
system. As will be shown in the Chapters dealing with mathematical modelling, the
structure of the net of characteristics is also very important in understanding the
numerical procedures required for the practical solution of hydraulic phenomena. In
other words, a good understanding of the physics of a hydraulic phenomena guides
us in the choice of appropriate numerical solution techniques and their numerical
parameters. As already discussed in § 3.4, the characteristics show us the initial- and
boundary data requirements for any hydraulic engineering problem simulation.

However, where the concept of characteristics is very important to us, the method of
characteristics provides us only for simple practical cases with an algorithm for the
computation of unsteady flow systems. For this reason we do not expand here on the
inclusion of energy generating or dissipating terms, such as gravity terms and
bottom friction, especially not for non-uniform channels with irregular topographies.
In this text we also do not show how to include lateral flows. Inclusion of such terms
is much easier and flexibly handled in programmable solution algorithms based on
implicit finite difference schemes, as discussed in Chapter 5.

We will return, however, to the practical use of characteristics in the case of water
hammer and flood propagation simulations.

WL | Delft Hydraulics 59
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

3.10 Questions and assignments

3.10.1 Derive Equations (3.1) and (3.3) again for the case where a channel
width b is introduced.

3.10.2 The simplified model is extended with a lateral flow term, by


introducing the lateral flow per unit width and length of the channel q
(m/s).
a. Derive the new form of Equations (3.1) and (3.3).
b. Give the transformation to the characteristic form.
c. Does this term give any change in the characteristic
directions?
d. Derive the new expressions for the Riemann invariants.

3.10.3 Calculate u and h at point P of Figure (3.5) for the following data:
Point A: u=0.5 m/s; h=3.00 m;
Point B: u=0.6 m/s; h=2.80 m.

3.10.4 A rectangular channel of 10 meter width carries a discharge of 6 m3/s.


For a water velocity u=2m/s, calculate the state of the flow. What
initial and boundary conditions do we need for modelling a certain
reach of the channel?

3.10.5 At the downstream end of a channel the flow is critical. A positive


characteristic arrives at this boundary carrying a Riemann invariant
J+=7 m/s. Calculate the outflow velocity and depth.

3.10.6. A gate in an irrigation canal of unit width is closed over a period of 4


minutes. The initial velocity of the water in the canal is 1.0 m/s at a
depth of 0.9 m. Assume a linear decrease of the discharge during the
closure.

a. Sketch the structure of the characteristics upstream and


downstream of the gate. Take a total time of 8 minutes and a
channel length of 2000 meter upstream and downstream of the
gate. Take one characteristic every 2 minutes (only for the purpose
of this exercise, as for practical solutions a somewhat denser net
will be needed).
b. Draw the hodograph plane and solve the depths and velocities at
the intersection points of the characteristics.
c. Draw a more precise network of characteristics.
d. Draw the water level and velocity functions along the canal at 8
minutes after the initiation of the closure.
e. What is going to happen at a later moment in time?

3.10.7. Repeat the same exercise for an instantaneous closure of the gate.

WL | Delft Hydraulics 60
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

4 Introducing Numerical Solutions for Partial


Differential Equations

4.1 Introduction
Partial differential equations are introduced when the dependent variables of
differential equations are a function of more than one independent variable. In many
practical applications these independent variables are x for space and t for time. In
this case, derivatives of the dependent variables have fractions both in space and in
time. For this reason we speak about partial derivatives, when dealing with a
derivative along one of these two or more independent variables. Consequently, we
speak about partial differential equations when rates of changes of these dependent
variables are related along the various dependent variables.

Depending of the nature of the physical processes, these partial differential


equations can be classified as belonging to one of the following types:

· hyperbolic equations;
· parabolic equations; and
· elliptic equations.

The classification of the Equation (s) for a certain problem leads to distinct
differences in which these equations are solved. Typical application areas for these
various classes are:

· hyperbolic equations: unsteady surface water flow; unsteady flow in the


unsaturated soil zone; flood wave propagation; transport of pollutants, channel
morphology;
· parabolic type: dispersion of pollutants; flood wave peak dampening; dispersion
of heat or energy;
· elliptic equations: steady surface water flow; steady ground-water flow.

The difference in type lies in the way in which disturbances propagate from one
point of a given (physical) system to other points. In the case of hyperbolic problems
it takes a definite time before a change of the state of the system at one point
influences the state at another point. The effect of these changes propagates along
so-called characteristics, or lines along which disturbances propagate. An obvious
example is the propagation of flood waves along rivers. Generally, it takes a few
hours or a number of days before a rain event in the upper catchments leads to
floods at locations downstream. In parabolic problems, the effect of a change
somewhere in the system, immediately affects the state everywhere else. In this case,
the characteristics travel at an infinite speed.

WL | Delft Hydraulics 61
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

In a step further, elliptic problems even have imaginary characteristics only and
these lead to an even more pronounced effect of the influence of the change of the
state at one single point on the state of the total system. Examples are state
descriptions of steady surface water or ground-water flow. It should be realised that
these are rather artificial descriptions, as true steady states do not exist in nature. For
this reason, most parabolic or elliptic problems are solved as if one is dealing with a
hyperbolic problem.

The simplest of the equations of hyperbolic type is the so-called advection equation
(see Equation (4.3)). Advection describes how some quantity, such as mass,
momentum, energy, heat, a pollutant etc. is transported by a carrier, such as water or
air. When dealing with the water sector, there is a variety of problems which can be
described by equations where advection processes play an important role. Examples
are:

· convective momentum term in the unsteady flow equations. Although often not
of much importance in most gradually varied flow situations, the advection term
(or convective momentum term) plays a dominant role in the formation of
hydraulic jumps, including the special appearances of moving hydraulic jumps,
such as tidal bores;
· flood wave propagation, as a special case of the application of unsteady flow
equations. Unsteady river- or channel flow is described by the De Saint Venant
equations. These form a set of second order partial differential equations of
hyperbolic type with two characteristic directions. In the most common case of
sub critical flow, these characteristics have opposite directions. In the special
case of flood wave propagation along rivers, gravity and friction play a dominant
role. This results in the emergence of an underlying simplified advection-
diffusion equation, of which the advection part is only first order hyperbolic. The
characteristic direction of this part is shown to us by the propagation celerity of
the flood wave peak;
· in hydrology also the movement of ground-water in the unsaturated zone is
based upon a combination of advection and diffusion processes, generally
described by the Richards equation. The advection part of this equation is driven
by gravity and by soil suction forces (matric forces). Due to the complex relation
between soil permeability and soil moisture, the process is extremely non-linear,
leading to the propagation of soil moisture shock fronts, much similar to the
propagation of hydraulic jumps in free surface flow;
· transport and dispersion of substances in water. A good description of these
processes is essential for the assessment of water quality processes, such as the
spreading of pollutants and heat, growth of algae, sedimentation processes etc.
An important element of this description, again, is the advection part of the
advection-dispersion equation.

WL | Delft Hydraulics 62
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

4.2 The advection equation


As an example, consider the advection of a conservative matter (e.g. a chemical
pollutant) in a channel. The derivation of the equation describing this process
proceeds via the introduction of a control element along the channel axis as shown
in Figure (4.1).

Figure 4.1 Control element for the derivation of the advection equation.

As a first step, the equation of mass conservation for water is derived. referring to
Figure (4.1), the mass balance for a control element with length dx along the x-axis
is set up by defining the change in storage to be equal to the difference between
inflow and outflow, or over a time period dt,

¶ ì é ¶ ü
( rV ) dt = í rQ - ê rQ + ( rQ ) dx ùú ý dt
¶t î ë ¶x ûþ

Replacing the volume V by the product of cross-sectional area A and distance step
dx it is found, for constant density , that

¶A ¶Q
+ =0 (4.1)
¶t ¶x

In a similar way, the mass balance for a conservative matter in the water with
concentration c can be derived as
¶ ¶
( Ac ) + ( Qc ) = 0 (4.2)
¶t ¶x

or, after differentiating out the terms of this equation and substituting Equation (4.1),

¶c ¶c
+u = 0 (4.3)
¶t ¶x

where the velocity u is defined as Q/A.

In this derivation it is assumed that the concentration c is uniformly distributed over


the cross-section. For the time being, it is also assumed that the cross-sectional area

WL | Delft Hydraulics 63
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

for flow is equal to the cross-sectional area representing storage. Equation (4.3) is
valid at every point of the x-t plane and shows the relation at the solution surface
between the tangent in x-direction (to the function which shows the distribution of c
along the river at a fixed time) and the tangent in t-direction (to the function showing
how c varies in time at a fixed point along the channel).

4.3 The characteristic solution


To get more insight into the meaning of Equation (4.3) let us consider the total
derivative

¶c ¶c
dc = dt + dx (4.4)
¶t ¶x

This equation simply states how the variation of c in any direction in the x-t plane is
composed of variations of c along the t- and x-axis, respectively. Division by dt
shows that, by equivalence of the Equations (4.3) and (4.4) along the line

dx
= u (4.5)
dt

the condition holds that

dc
= 0 (4.6)
dt

or, after integration of Equation (4.6), that c remains constant along this line. The
line defined by Equation (4.5) is called the characteristic of Equation (4.3). In
Figure (4.2) it is shown how a solution of Equation (4.3) is obtained, for the simple
case where u=constant and where, consequently, the characteristics are parallel lines
in the x-t plane.

Figure 4.2 Solution of the advection equation with the method of characteristics.
It is also seen that a solution is only obtained when in the computational domain
initial data are defined along the line t=0 and, where an upstream boundary is

WL | Delft Hydraulics 64
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

present, also boundary data along the line x=0. As will be shown later, these
characteristics play an important role in defining where initial- and boundary data
are required, as well as the number and the type of these conditions. For practical
applications the method of characteristics is not often applied in this form, mainly
because it has the disadvantage that it is not based upon a fixed grid in space.
This makes it more complicated to include in the solution such processes as sources
and sinks, decay and diffusion, which will be discussed later.

4.4 Finite difference schemes


On the grid, shown in Figure (4.3), the derivative c/ x can be written in several
ways in terms of finite differences and, at grid point (i,n) more in particular as

¶c cin - cin-1
@ backward difference approximation;
¶x Dx
¶c ci +1 - cin-1
n
@ centred difference approximation;
¶x 2Dx
¶c ci +1 - cin
n
@ forward difference approximation.
¶x Dx

Figure 4.3 Grid for defining finite difference schemes of the advection equation.

In a similar way finite difference approximations for c/ t can be defined. For the
advection scheme these approximations can be combined to nine possible finite
difference schemes, already. This number even increases when combinations of 1),
2) and 3) are made by introducing, for example, other time levels than n t. Not all
these schemes will give satisfactory results and it is needed, therefore, to define
criteria for judging which scheme can be selected as the best out of the large number
of possible schemes.

Let us consider, in the first instance, a scheme approximated by a backward


difference in the x-direction and a forward difference in the t-direction. Equation
(4.3) than reads as

WL | Delft Hydraulics 65
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

cin +1 - cin c n - cin-1


+u i = 0
Dt Dx

or, expressing cin+1 in terms of the other variables and defining


uDt
r = (4.7)
Dx

cin +1 = r cin-1 + (1 - r ) cin (4.8)

This relation can be represented graphically by the operator shown in Figure (4.4).

Figure 4.4 Operator for the scheme of Equation Error! Reference source not found..

A solution using this scheme is shown in Figure (4.5) for the following data:

u = 0.0185 m/s;
x = 100 m;
t = 0.75 hours.

Upstream boundary data are set as c=0 and the initial data are given by the Gaussian
distribution of the form
x-m 2
-½ ( )
c = e s
(4.9)

with mean value =500 m and standard deviation =100 m.

WL | Delft Hydraulics 66
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 4.5 Numerical solution of the advection equation using the scheme given by Equation (4.8)
with r=½ and various values for x.

Figure (4.5) shows the effect of varying numerical parameters on the solution. The
exact solution is defined by a shift of the initial Gaussian distribution over 400 m
along the x-axis. Numerically, the same solution is found for r=l. For r=½ the wave
is strongly dampened. This effect is called numerical diffusion of the wave. The
figure also shows that a smaller value for x, while maintaining the same value r=½,
reduces the numerical diffusion. For a value r=2 the solution will not look realistic.
Negative values for c are found as well as values greater than 1. These values cannot
possibly be correct as the method of characteristics shows that in the absence of
source- and sink terms extreme values for c are found along the boundaries where
characteristics enter the computational domain. The problems faced here are related
to the stability of the scheme.

It is clear that not only the choice of a specific scheme is important for obtaining
good results in a computation, but also the choice of the numerical parameters, such
as x and t, for that scheme. The accuracy aspects of a particular scheme can be
judged by writing each term in a Taylor's series expansion from the selected centre
point of the scheme. At the centre point (i,n) we obtain for the scheme given by
Equation (4.8):

n n n
n +1 Dt æ ¶c ö Dt 2 æ ¶ 2 c ö Dt 3 æ ¶ 3 c ö
c i = c + n
i ç ÷ + ç ÷ + ç ÷ + h.o.t .
1! è ¶t øi 2! è ¶t 2 øi 3! è ¶t 3 ø i

and

= c +
-Dx æ ¶c ö
n
Dx 2 æ ¶ 2 c ö ( -Dx 3 ) æ ¶ 3c ö
n n

ç ÷ + ÷ + ÷ + h.o.t .
n n
c i -1 i ç ç
1! è ¶x øi 2! è ¶x 2 øi 3! è ¶x3 øi

WL | Delft Hydraulics 67
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where h.o.t. stands for higher order terms. Collecting terms in Equation (4.8) and
dividing by t gives at point (i,n):

¶c ¶c -Dt ¶ 2 c r Dx 2 ¶ 2c Dt 2 ¶ 3c r Dx 3 ¶ 3c
+u = + - - + h.o.t. (4.10)
¶t ¶x 2 ¶t 2 2Dt ¶x 2 6 ¶t 3 6Dt ¶x 3

The right hand side of this equation is called the truncation error of the finite
difference scheme. It can be concluded that by using the scheme of Equation (4.8),
another equation, containing higher order derivatives, is solved than the differential
equation (4.3). For looking more closely at the lowest order terms of the truncation
error, Equation (4.3) is differentiated with respect to t to give, for constant u,

¶ 2c ¶ 2c ¶ 2c ¶ 2c
= -u = -u = u2
¶t 2 ¶x¶t ¶t ¶x ¶x 2

Substitution of this relation into Equation (4.10) gives

¶c ¶c ¶ 2c
+u = ½u Dx (1 - r ) + h.o.t. (4.11)
¶t ¶x ¶x 2

Referring already to Equation (4.17), it will be shown in Chapter 4 that this equation
adds diffusion to the advection equation. For the example of Figure (4.5), with r = ½
and x = 100 m, the numerical diffusion coefficient is 0.463 m2/s. Comparison of
Figure (4.5) with Figure (4.9) of § 4.7 confirms the magnitude of the diffusive effect
of the truncation error.

Equation (4.11) shows, furthermore, that the lowest order term of the truncation
error goes to zero for r=1. The computed results for r = l indicate that a further
substitution will cancel out all the higher order terms in the truncation error for this
specific value of r, to give the exact solution of the differential equation. For values r
= ½ the numerical diffusion can be reduced by taking smaller and smaller values for
x. In the limit, for x --> 0, the truncation error will go to zero (see Equation
(4.11)) so that the numerical solution converges towards the true solution of
Equation (4.3). This convergence will not be found for the solutions with r=2. It can
be shown that for this r-value subsequent reduction of x will give a solution
deviating more and more from the true solution. The results become more and more
unstable. There are several ways in which this stability problem can be investigated:

1. by interpreting the dominant effects of the truncation error of the scheme;


2. by showing that the values in the computational domain will always remain
between the extreme values specified at the boundaries;
3. by an analysis based on the role of characteristics on the solution (Courant-
Friedrich-Lewy or CFL-condition);
4. by a Fourier series analysis of the solution function and its transformation
from one time level to another.

Ad 1) For positive values of the numerical diffusion coefficient in Equation (4.10),


the solution will be dampened within the computational domain.

WL | Delft Hydraulics 68
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Negative diffusion coefficients are physically unrealistic and would lead to


amplification of the solution. This would affect most strongly the shorter wave
lengths as those have the largest values for 2c/ x2. In this way, small errors will
rapidly be blown up to large values. This leads to the conclusion that only positive
diffusion coefficients will give stable results and thus, from Equation (4.11), that r
should not exceed unity.

Ad 2) The exact solution of Equation (4.3) shows that extremes in the computational
domain should always remain between minimum and maximum values given at the
boundaries of this domain. It can be shown that the scheme, indeed, gives values
which are limited by the extremes at the boundaries for r 1 (Abbott, 1979).

Ad 3) Referring to Figure (4.7) and following the characteristics through point


(i,n+l), the scheme will be stable when the characteristic intersects the line t=n t
between the grid points (i-1,n) and (i,n), as in this case cin+1 is found from an
interpolation of earlier computed results. When the intersection of the characteristic
and the line t=n t is outside the line element between the points (i-l,n) and (i,n), the
value cin+1 follows from an extrapolation of data and the scheme will become
unstable. The time step can be compared with the time tc required to travel along
the characteristic over a distance x. The ratio t/ tc is defined as the Courant
number Cr. For Cr 1 the scheme will be stable and it will be unstable for Cr>1. This
case is shown in Figure (4.7) when, for the characteristic leaving the point (i-1,n+1)
backward in time, the numerical scheme of Equation (4.8) is applied at the grid point
(i-1,n).

Ad 4) A solution at time level n t is decomposed into the Fourier series


x j
ik p ik p
c n
j = åxk
n
k e l
= åx
k
n
k e jj
(4.12)

where
im = indicator for imaginary numbers;
k = wave component number;
n
k = wave amplitude at time level n t;
L = length of the computational domain along x;
ii = highest grid point number over domain L.

Stability of the solution is guaranteed if, in the transformation of the solution from
time level n t to time level (n+l) t by the scheme, none of the wave amplitudes kn
are amplified. Substitution of Equation (4.12) into Equation (4.8) gives

i
im k p ì im k i -1p im k p ü
i
x kn +1 e ii
= x kn íre ii + (1 - r ) e ii ý
î þ
or
x kn +1 k
im p
Ak = = 1- r + e ii
(4.13)
x kn

WL | Delft Hydraulics 69
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

For a specific wave component the amplification |Ak| is defined as the length of the
vector composed of the real and imaginary parts of Equation (4.13). From Figure
(4.6) it can be concluded that |Ak| 1 for r 1.

Figure 4.6 Argand wave amplification diagram for the scheme of Equation (4.8)

Table 4.1 Example of unstable solution of Equation (4.8) for r=2

i => 0 1 2 3 4 5 6 7 8 9 10
x (m) => 0 100 200 300 400 500 600 700 800 900 1000
n t (s)
0 0 0 1 0 0 0 0 0 0 0 0 0
1 200 0 -1 2 0 0 0 0 0 0 0 0
2 400 0 1 -4 4 0 0 0 0 0 0 0
3 600 0 -1 6 -12 8 0 0 0 0 0 0
4 800 0 1 -8 24 -32 16 0 0 0 0 0
5 1000 0 -1 10 -40 80 -80 32 0 0 0 0
6 1200 0 1 -12 60 -160 240 -192 64 0 0 0
7 1400 0 -1 14 -84 280 -560 672 -448 128 0 0
8 1600 0 1 -16 112 -448 1120 -1792 1792 -1024 256 0
9 1800 0 -1 18 -144 672 -2016 4032 -5376 4608 -2304 512
10 2000 0 1 -20 180 -960 3360 -8064 #### #### #### -5120
11 2200 0 -1 22 -220 1320 -5280 #### #### #### #### ####
12 2400 0 1 -24 264 -1760 7920 #### #### #### #### ####
13 2600 0 -1 26 -312 2288 #### #### #### #### #### ####
14 2800 0 1 -28 364 -2912 #### #### #### #### #### ####
15 3000 0 -1 30 -420 3640 #### #### #### #### #### ####

4.5 Characteristic solutions on a fixed grid


The accuracy of the scheme given by Equation (4.8) is rather low for many practical
applications. It may be possible to choose r-values equal or close to unity on some
parts of the grid, but varying velocities may make it necessary to work with lower
values elsewhere. Reduction of x increases rapidly the number of operations and
may demand much computer time.

WL | Delft Hydraulics 70
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Higher accuracy schemes introduce the need to define relations between more grid
points. Programming wise this may lead to a rather complicated organisation. For a
relatively clear and accurate procedure we will return now to the method of
characteristics. For the fixed grid, as shown in Figure (4.7), one may follow the
characteristic line from grid point (i,n+l) back to the time level n t.

Figure 4.7. Characteristic solution on a fixed grid.

As the value for c remains constant along this characteristic, the accuracy of the
solution technique depends on the accurate positioning of the intersection point at
time level n t and on an accurate interpolation. For a constant velocity u and a
linear interpolation between c-values at points (i-l,n) and (i,n) it is found that

uDt n Dx - u Dt n
cin +1 = ci -1 + ci
Dx Dx

This relation is equivalent to the finite difference scheme given by Equation (4.8).
The advantage of this method is that it can be extended to Courant numbers greater
than unity (r>1; see Figure (4.7). An improvement of the accuracy is obtained by
using a cubic spline function interpolation (Preissmann and Holly, 1977). In this
interpolation not only the values for c are used at the grid points (i-l,n) and (i,n), but
also their first derivatives in x. The interpolation is than given by
n n
æ ¶c ö æ ¶c ö
c n +1
i = ac n
1 i -1 + a c + a3 ç ÷ + a 4 ç ÷
n
2 i (4.14)
è ¶x øi -1 è ¶x øi
with
a1 = r 2 (3 - 2r )
a 2 = 1 - a1
(4.15)
a 3 = r 2 (1 - r )Dx
a 4 = - r (1 - r ) 2 Dx

Preissmann and Holly describe the computation of c/ x values via an advection


scheme. As under simplified conditions the advection does not alter the shape of the
translated function, it must be possible to find an equation for the advection of both
the first and the higher order derivatives. For the fractioned step approaches
described later, this procedure, however, cannot be followed. A simpler, but also less
accurate procedure, is found by approximating the derivatives from earlier computed
c-values as

WL | Delft Hydraulics 71
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

¶c cin+1 - cin-1
@ (4.16)
¶x 2Dx

The results obtained with this method of characteristics, using cubic spline functions
for interpolation, show a rather low numerical diffusion. However, some parasitic
waves still appear in the tail of the function.

4.6 Introducing diffusion


The derivation of the equation for advection-diffusion processes proceeds very
similar to the one for advection alone. Referring again to Figure (4.1), the advection
through the element boundary has to be increased with the contribution of molecular
and turbulent exchange processes that can be described by Fick's law as

¶c
Tdispersion = - DA
¶x

where D is a diffusion coefficient. The balance equation now reads

¶ ¶ ¶ ¶c
( Ac ) + ( Qc ) = - æç - DA ö÷
¶t ¶x ¶x è ¶x ø

or, assuming that the effects of spatial variation in cross-section and diffusion
coefficients can be neglected,

¶c ¶c ¶ 2c
+u = D (4.17)
¶t ¶x ¶x 2

The contribution of molecular diffusion to D is negligible in practical applications.


Furthermore, in a schematization where u and t are averaged over the cross-section,
the value of D is much larger than following from turbulent diffusion alone. Much of
the spreading of the matter along the channel axis is caused by differential
advection. Differences in advection are caused by velocity variations in the vertical
profile, in the direction perpendicular to the flow lines, as a result of spiral flow in
bends, due to storage elements along the channel banks etc. The combined effect of
turbulent diffusion and differential advection is called dispersion, so that D in
Equation (4.17) generally stands for a dispersion coefficient.

Equation (4.17) has an exact solution defined as

( x -ut )2
M0 -
c ( x, t ) = e 4 Dt
(4.18)
2 A (p Dt )
½

where M0 is the total mass of a tracer which is injected instantaneously and


uniformly distributed over the complete cross-section. The initial distribution is

WL | Delft Hydraulics 72
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

defined as a Dirac delta distribution, which tends to be dispersed along the channel
axis as a Gaussian distribution with mean value =x+ut and with a standard
deviation = 4Dt .

For an initial Gaussian distribution the value of 0 can be transformed into a value,
defined as the time which it would have taken to come from a delta distribution to
the given Gaussian form. This leads to
s 02
t0 = (4.19)
4D

and the exact solution defined as

{x -( x0 +ut )}
2

-
M0 4 D ( t0 + t )
c ( x, t ) = e (4.20)
2 A {p D ( t0 + t )}
½

where M0, in this case, is the total mass of the tracer integrated along the channel
section at time t=0.

4.7 An explicit finite difference scheme


Exact solutions are defined only for very particular cases. The following limitations
of the applicability can be mentioned:

· the solution is defined only for single point instantaneous sources ;


· the channel characteristics are often not constant along x.

More general, again, practical solutions are found with finite difference
approximations, which will be demonstrate here, in the first instance, on a scheme
for the diffusion part of the equation only, given by

¶c ¶ 2c
= D (4.21)
¶t ¶x 2

Consider a grid, as shown in Figure (3.8).

Figure 4.8 Operator for an explicit scheme for the diffusion equation.

WL | Delft Hydraulics 73
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The time derivative can be given as a forward finite difference approximation in the
form

¶c cin +1 - cin
@
¶t Dt

The second derivative in x is approximated as

n n
æ ¶c ö æ ¶c ö
ç ÷ -ç ÷
¶ 2c ¶ æ ¶c ö è ¶x øi +½ è ¶x øi -½
= ç ÷ @ @
¶x 2 ¶x è ¶x ø Dx
æ cin+1 - cin ö æ cin - cin-1 ö
ç ÷-ç ÷
è Dx ø è Dx ø =
cin-1 - 2cin + cin+1
Dx Dx 2

Substitution of these finite difference approximations into Equation (4.21) gives the
scheme, with cjn+1 expressed in terms of variables at time level n t,

cin +1 = rcin-1 + (1 - 2r )cin + rcin+1 (4.22)

where
Dt
r = D (4.23)
Dx 2

As the value at grid point (i,n+l) is computed explicitly from known values at time
level n t, it is customary to call this an explicit scheme. At a later stage we will
introduce a somewhat refined definition for the types of finite difference schemes. A
Taylor's series development shows that, in reality, we are solving with the finite
difference scheme (4.22) the equation

¶c ¶ 2c -Dt ¶ 2 c 1 2 ¶ c
4
-D 2 = + D Dx + h.o.t.
¶t ¶x 2 ¶t 2 12 ¶x 4

An analysis similar to the one for the advection equation leads to the cancellation of
the lowest order terms in the truncation error for r=1/6. However, for this same value
of r, cancellation of the higher order terms does not occur. A complete analysis will
lead to the conclusion that, for this scheme, it is impossible to find any r-value which
will produce the exact solution.

As an example, consider a channel section of length L, where initially a conservative


pollutant with concentration c is present, given as a Gaussian distribution with
standard deviation x. Selecting a distance step x=100 m and a dispersion
coefficient D=0.463 m2/s, a time step t=l hour will give an r-value 1/6.

WL | Delft Hydraulics 74
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 4.9 Results with the numerical scheme of Equation (4.22) for various values of r.

Consistent with the theory of partial differential equations, the solution requires one-
point initial data at t=0 and one point boundary data at x=0 and x=L. In the case of
this example, constant values c=0 have been set at both boundaries. For the second
order space derivative of this parabolic partial differential equation, both boundary
data could also have been given as first order derivatives of c along x.

In Figure (4.9), a comparison is shown between the exact solution for this problem
and numerical solutions obtained for different grid steps. With respect to the choice
of the grid steps it is interesting to note that, unlike for the advection scheme, the
highest accuracy is not found at the limit of stability. For practical applications it is
best to choose a grid which gives r-values around 1/6. However, Figure (4.9) also
shows that the differences are minor. This is certainly true when compared with the
sensitivity of numerical parameters in advection schemes. In all cases, it is important
to make sure that the stability limit r=½ is never exceeded. The existance of such
limit can be investigated again with a Fourier series analysis, to give an
amplification factor

kmDx
A = 1 - 4r sin 2 (4.24)
2

so that, since the sin2-term is definite positive, the stability condition |A| 1 implies
that

r £ ½ (4.25)
or
Dx 2
Dt £ (4.26)
2D

WL | Delft Hydraulics 75
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

4.8 Explicit schemes for the combination of advection and


diffusion
For a first discussion, let us return to the advection scheme given by Equation (4.8).
Comparing the Taylor's series expansion of this scheme shown by Equation (4.11),
to the advection-diffusion Equation (4.17) it is seen that physical and numerical
diffusion are equal when

D = ½u Dx (1 - r ) (4.27)

In principle, advection-diffusion processes could be computed with a scheme for the


advection equation alone, by selecting numerical parameters which just give the
correct amount of diffusion by satisfying Equation (4.27). In practice this cannot
easily be realised as u varies in general with x and t and the grid steps cannot be
adjusted from one to the other grid point. Furthermore, the numerical diffusion
coefficient has been derived from the idealized case where u and D are constant. It
is, therefore, more practical to combine a diffusion scheme with a scheme for
advection, which has a numerical diffusion which is small as compared to the
physical diffusion.
A convenient approach is the fractioned step method where the contributions to the
variations in c by advection and dispersion are solved sequentially. As a first step,
the equation

¶c ¶c
+u = 0 (4.28)
¶t advection ¶x

is solved, so as to give the c-values at (n+l) t resulting from transport alone, while
as a second step the equation

æ ¶c ö ¶c æ ¶c ö ¶ 2c
ç ÷ = -ç ÷ = D (4.29)
è ¶t ø dispersion ¶t è ¶t ø advection ¶x 2

is solved, giving the complete solution for advection and dispersion together.
Equation (4.28) can, again, be solved with the fixed grid characteristic method using
the spline function interpolation (see Equation (4.14)), giving intermediate values cj*
defined by advection alone. Subsequently, the dispersion step can be given in the
finite difference form as

cin +1 - ci* ci*-1 - 2ci* + ci*+1


@ D
Dt Dx 2

or

cin +1 = rci*-1 + (1 - 2r )ci* + rci*+1 (4.30)

WL | Delft Hydraulics 76
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The advection-diffusion equation can easily be extended to the two-dimensional


case, giving

¶c ¶c ¶c ¶ 2c ¶ 2c
+u +v = Dx + D y (4.31)
¶t ¶x ¶y ¶x 2 ¶y 2

where y is the direction in the horizontal plane orthogonal to x, v is the water


velocity in y-direction and Dx and Dy are dispersion coefficients in x- and y-
direction, respectively. This equation can be solved, following the same principles,
in the sequence of fractioned steps:

1) advection along x;
2) advection along y;
3) dispersion along x;
4) dispersion along y.

4.9 Implicit schemes for the diffusion equation


The stability criterion for the diffusion scheme of Equation (4.22) limits the time
step of the solution method and this is not always desirable. For this reason implicit
schemes can be introduced, which have less stringent stability conditions. As shown
in Figure (4.10), in implicit schemes the unknown variables at the new time level
(n+1) t are connected by the partial definition of the space derivative of Equation
(4.21) at the new time level as

cin +1 - cin
Dt
@
D
Dx 2
{ }
(1 - q ) ( cin-1 - 2cin + cin+1 ) + q ( cin-+11 - 2cin+1 + cin++11 ) (4.32)

where defines a weighting of the space derivative between the time levels n t and
(n+1) t. This equation can be rewritten to the form

a i cin-+11 + bi cin +1 + g i cin-+11 = d i (4.33)

where
ai = - q r
bi = 1 + 2q r
(4.34)
g i = -q r
di = (1 - q ) r cin-1 + (1 - 2r (1 - q ) ) cin + (1 - q ) r cin+1
As an alternative, the implicit finite difference scheme may be written in terms of
changes c from time level n t to (n+1) t as

Dci
Dt
@
D
Dx 2
{( c
n
i -1 }
- 2cin + cin+1 ) + q ( Dci -1 - 2Dci + Dci +1 ) (4.35)

which can be written in a form similar to that of Equation (4.33).

WL | Delft Hydraulics 77
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 4.10 Grid for the implicit scheme of the diffusion equation

The system of Equations (4.33) can be solved with the double sweep algorithm, a
special form of the Gauss elimination, based upon tri-diagonal materices. In this
algorithm, in a first sweep the coefficients in the lower diagonal are eliminated,
starting with the left hand boundary condition. In a second sweep values for the
unknows are computed, by subsequently eliminating the upper diagonal coefficients.

Figure 4.11 Schematic presentation of the tri-diagonal double sweep matrix solution algorithm

Schematically, the algorithm is presented in Figure (4.10). Assuming the left hand
and right hand boundary conditions, respectively, to be of the form

b 0 c0 + g 0 c1 = d0 (4.36)
a ii cii -1 + bii cii = d ii (4.37)

Gaussian elimination of the lower diagonal leads to the set of equations

cin +1 + g i*cin-+11 = d i* (4.38)

with the new coefficients in the matrix (now marked with a superscript *), computed
from grid point 1 until ii-1 by the relations

g d i - ad i*-1
g i* = ; d i* = (4.39)
b - ag i*-1 b - ag i*-1

WL | Delft Hydraulics 78
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

For the more general case of non-linear equations, with coefficients derived as a
function of x, also the coeficients , and in Equation (4.33) will get a subscript i.
These subscripts will than also appear in the relations given by Equation (4.39).

The elimination is started by assigning the starting values for the new coefficients at
grid point 0 by substitution of the boundary relation given by Equation (4.36) into
Equation (4.33) applied at gid point 1, giving

g0 d0
g 0* = ; d 0* = (4.40)
b0 b0

The same result would also have been obtained simply by setting the value of 0 to
zero in Equation (4.39). In this case the assignment of new values for i* and i*
could have started at grid point 0 instead of 1.

After completion of the first sweep, the right hand boundary value is obtained by
substitution of the boundary relation given by Equation (4.37) into Equation (4.39)
at point ii-1. Successive back substitution (back sweep) of the values for c leads to
the overall solution of the system of equations.

WL | Delft Hydraulics 79
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

WL | Delft Hydraulics 80
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

5 De Saint Venant equations and their


solutions

5.1 Introduction
In Chapter 3 the unsteady channel flow equations were introduced for canals with
simple rectangular cross-sections. These equations will now be extended to the more
general case of irregular cross-sections. In this approach, the flow in rivers and other
type of channels is assumed to be correctly defined by the state variables or
dependent variables discharge, Q and water level, as a function of the independent
variables t for time and x for space.

Basic assumptions and/or limiting conditions are:

· the discharge is sufficiently well defined as the integral of the velocities through
a cross-section, perpendicular to the x-axis and perpendicular to the flow
velocity vectors in the flood plain;
· the water level is constant along the cross-section. This implies that at any time
the water level at all points along a given cross-section should rise or fall at the
same rate. This assumption is generally justified when the widths of river and
flood plain are of the same scale and free of obstacles such as natural levees or
embankments;
· the water level slope or gradient in the x-direction is constant along the cross-
section. However, it should be noted that in case this condition is not satisfied,
correction factors may be applied to reduce significantly errors in the model
parameters, as discussed in this contribution.

Quantitative analysis of the two state variables requires two independent equations.
Usually, the following equations are used:

· the continuity equation, based upon volume conservation in a control volume


defined between two successive points along the channel axis;
· the momentum equation, based upon the conservation of momentum, including
the effect of impulses generated by forces acting upon the water contained in the
control volume.

Making, furthermore, the following assumptions:

· the pressure distribution in the vertical is hydrostatic;


· the resistance relationship for steady flow is also applicable for unsteady flow;
and
· the bed slope is moderately steep, so that the cosine of the slope can be replaced
by unity,

WL | Delft Hydraulics 81
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

De Saint Venant (1871) derived the following equations (presented in a slightly


adapted form here):

¶As ¶Q
+ = ql (5.1)
¶t ¶x

1 ¶Q ¶ Q 2 ¶z Q Q
{ + ( ) }+ + 2 = 0 (5.2)
gA ¶t ¶x A ¶x K

where As is the cross-sectional area representative for storage over a control volume
(m2), t the time (s), Q the discharge (m3/s), x the position along the channel axis (m),
ql the lateral discharge per unit length of channel (m2/s), A the flow conveying cross-
sectional area (m2), the stage or water level above a selected horizontal reference
plane (m) and K the channel conveyance (m3/s). The meaning of the parameters in
these equations becomes clear when we follow the derivation of these equations.

5.2 The continuity equation


The continuity equation for channels with arbitrary cross-section can be derived on
the basis of the same principles as applied to the case of the simplified rectangular
profile, as described in Chapter 3. It should be realised that in channels and rivers
with arbitrary cross-sections even the topography between two successive cross-
sections is irregular. For this reason, we will base the derivation of the continuity
equation upon a storage volume V( ), as a function of stage .

Figure 5.1 Control volume for the derivation of the continuity equation

From Figure (5.1) it is readily seen that the following volume conservation equation
holds:

¶V ¶Q
+ dx = ql dx
¶t ¶x

As for each control volume between two successive points along x, the volume is a
function of the water level, the following substitution can be made

WL | Delft Hydraulics 82
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

¶V dV ¶V ¶V ¶V
= = Asurf = b s dx (5.3)
¶t d V ¶t ¶t ¶t

where Asurf is the storage area in the horizontal plane defined between two
successive control sections. It is common practice, however, to define the storage as
a cross-section property. For this reason, we usually transform the parameter Asurf to
a parameter linked to a cross-section, by introducing a storage width bs of the
channel. This level dependent channel cross-section storage parameter bs is defined
as Asurf (for a given level) divided by the distance between two successive control
sections (Equation (5.3)). This definition leads to the continuity equation of the
commonly used form

¶V ¶Q
bs + = ql (5.4)
¶t ¶x

Integration of bs over the height of the cross-section leads to the definition of the
storage cross-section
V
As (V ) = ò b dzs (5.5)
zb

as introduced earlier in Equation (5.1). It should be realised that the storage cross-
section parameters bs and As are often different from the parameters B and A used in
the momentum equation (see § 5.3). These differences will be discussed in more
detail in § 5.7.

The second important parameter in the continuity equation is the lateral flow.
Examples of lateral flow contributions, either positive or negative, are:

· catchment runoff hydrographs, either given as point sources or as distributed


sources. These consists both of surface water and ground-water components;
· drainage releases, for example, via pumps;
· irrigation water extraction;
· flow over side weirs;
· water exchange at the river bed, possible linked to a ground-water storage
description;
· rainfall and evapotranspiration directly linked to the water surface.

WL | Delft Hydraulics 83
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

5.3 The momentum equation


For the derivation of the momentum equation we will limit ourselves here to a
pragmatic derivation. Recalling § 3.3, with the derivation of the momentum equation
for a uniform rectangular cross-section, one of the intermediate forms has been

¶ ¶ ¶h
(uh) + ( u 2h) + gh = 0 (5.6)
¶t ¶x ¶x

Referring to Figure (5.2), the momentum equation of the overall cross-section is


obtained by integrating each term of Equation over the width.

Figure 5.2 Cross-section for the integration of the momentum equation

Integration of the first component leads to the acceleration term


B B
¶ ¶ ¶Q
òo ¶t ( u j h j )dy = ¶t òo
u j h j dy =
¶t

Integration of the second component leads to the convective momentum term


B B 2
¶ ¶ ¶ ¶ Q
òo ¶x ( u i hi ) dy = ò ( u 2j h j ) dy = ( b u Q) = (b
2
)
¶x o ¶x ¶x A

where is the Boussinesq coefficient of velocity distribution. It corrects for the fact
that we have used the average flow velocity u in the integration of the momentum
carried by the flow through the cross-section, instead of the local velocities.
Consequently, the value of is greater than unity. It should be noted that the
importance of this convective momentum term is generally limited and in most
practical applications the value of ß is simply set to unity. In some practical
applications the value of ß is even set to zero, which means that the complete term is
neglected (see diffusive wave approximation of Chapter 6).

WL | Delft Hydraulics 84
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 5.3 Definition of the gravity component

Integration of the second component leads to the combined hydrostatic pressure and
gravity term
B B
¶V ¶V ¶V
òO g h j ¶x dy = g ¶x ò h dy
O
j = gA
¶x

In the integration it has been assumed that the term (zb+h)/ x, or x, is constant
across the section. In order to separate the effects of hydrostatic pressure gradients
and gravity, the term can be split up again on the basis of a kind of representative
depth hr above a representative bottom level zbr as follows

¶V æ ¶z ¶h ö æ ¶h ö
gA = gA ç br + r ÷ = gA ç I 0 r + r ÷ (5.7)
¶x è ¶x ¶x ø è ¶x ø

It can be concluded from Equation (5.7) that the way in which we define this bottom
level does not affect the results of computations. This definition is a product of our
mind and not that of the river itself. The river feels for its flow only the effect of the
water level slope (combined with cross-sectional parameters) and not that of a bed
slope, as such.

Finally, a friction term has to be added. This friction description is rather empirical
and it is common practice to follow the same definition as applying to steady flow.

In steady flow, the bottom friction for sub section j of the cross-section is supposed
to balance the gravity term, or

( t bottom) j = r g h j I o

Furthermore, t =t (u2) for turbulent flow, leading to the Chezy relation

1/ 2
u j = Cj h j Io

or to the Manning relation

WL | Delft Hydraulics 85
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

1
u j= hj2/3 I0
nj

Proceeding with the Manning relationship and taking into account the side slope of
the bottom profile, the local depth is generally replaced by a local hydraulic radius
Rj, giving

1
u j= R j 2/3 I0
nj

For unsteady flow, the slope expressed by this shear force relation deviates from the
bottom slope and for each sub section a friction slope may be introduced defined as

2
u 2j A2j Qj
If = 2
= 2
æ1 2/3
ö Kj
çç Aj R j ÷÷
è nj ø

where Kj is the channel conveyance for subsection j.

As per definition of a one-dimensional schematisation If is constant over the cross-


section, integration of the friction term over this cross-section leads to a contribution
to the momentum equation is given as

2
Q
If = 2
(5.8)
K

with the total channel conveyance K defined as

jj
1
K = ån
j =1 j
Aj R j 2 / 3 (5.9)

where jj is the number of sub sections in the channel cross-section.

The concept of the total conveyance of a cross-section is based on the integration or


summation of the contributions Kj for individual vertical slices of the cross-section.
For each individual slice (Figure 5.2) the concept of the hydraulic radius is used
where
Aj
Rj= (5.10)
Pj

with the wet perimeter Pj representing the contact length of water and channel bed.

WL | Delft Hydraulics 86
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

It should be remarked that the concept of a wet perimeter is based on the assumption
that the shear force is equal at all points along this perimeter. This is approximately
true for a small sub-section or slice of the channel cross-section. However, it is not
true for the complete cross-section, as the local shear force is a linear function of the
local depth. It is, therefore, strongly advised not to use the concept of the hydraulic
radius for rather irregular cross-sections. For such sections one should always work
with the concept of the conveyance, or discharge capacity K, obtained through the
integration of the local conveyance across the section.

In the special cases where the hydraulic radius concept can be used for the oveall
cross-section, the Manning and Chezy coefficients are related through

1
Q = K I = AR 2 / 3 I (5.11)
n
and

Q = K I = CAR 1/ 2 I (5.12)

giving the relation between the Manning and Chezy coefficients as

1 1/ 6 1 1/ 6
C = R ; or n = R (5.13)
n C

It is quite well accepted that the Manning coefficient is more constant with depth
than the Chezy coefficient. Especially at relatively low flow depth, such as occurring
on flood plains, it is recommended to use the Manning coefficient for the friction
description.

The terms of the momentum equation can now be collected to give the form

2
¶Q ¶ Q ¶h Q|Q|
+ (b ) + gA r + gAI 0 r + gA = 0 (5.14)
¶t ¶x A ¶x K
2

Division of all terms by gA gives the equation in dimensionless form as

1 ¶Q ¶ Q 2 ¶h QQ
{ + ( ) } + r + I0r + 2 = 0 (5.15)
gA ¶t ¶x A ¶x 1424 K3
144 42444 3
kinematic wave

14444444 4244444444
diffusive wave
3
dynamic wave

which is equivalent to Equation (5.2). Equation (5.12) also shows how the
momentum term can be simplified when some parts of this equation are of lesser
importance. These kinematic wave and diffusive wave approximations will be
discussed in § 6.4.

WL | Delft Hydraulics 87
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

5.4 Numerical solutions


Hydraulic modelling techniques for 1D are currently based upon the numerical
solution of the de Saint Venant equations, including the full convective momentum
term. These numerical solutions are nearly exclusively based upon implicit finite
difference methods. These offer the advantage of unconditional numerical stability,
while the various robustness problems of the past, relating to non-linear effects and
flooding and drying of channels and flood plains, have been solved satisfactorily.
The application of finite elements and finite volume techniques does not provide
specific advantages as, in 1D modelling, most of these techniques lead to equivalent
forms derived through finite difference formulations.

Numerical methods for the de Saint Venant equations may be based upon so-called
staggered and non-staggered schemes. The first category represents formulations
where the dependent variables Q and are defined alternatingly at successive grid
points along the x-axis. For non-staggered schemes, however, the variables Q and
are defined at the same grid points. At first sight this last definition offers
advantages through the availability of the state variables discharge and water level at
the same points along the channel axis. It has been shown, however, that the
staggered grid approach offers distinct advantages over non-staggered grids by
guaranteeing the convergence of numerical solutions and the better ability to handle
flooding and drying of grid sections, as shown by Stelling et al. (1998).

For the numerical solution of the de Saint Venant Equations (5.1) and (5.2), we will
consider their Eulerian form per unit width of channel by first neglecting the lateral
flow and further simplifying the equations to

¶V ¶ (uh)
+ =0 (5.16)
¶t ¶x
and
¶u ¶u ¶V uu
+u +g + cf =0 (5.17)
¶t ¶x ¶x h

where is the water level defined as = h+zb with h defined as the local water depth
(m) and zb as the local bottom level (m), u the flow velocity (m/s) and cf the
dimensionless bottom friction coefficient.

Figure 5.4 Staggered grid for unsteady channel flow

WL | Delft Hydraulics 88
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Referring to Figure 1 and to Stelling and Duinmeijer (2003) for further details, the
staggered grid approach requires that alternatingly at - and u-points, Equations
(5.16) and (5.17) are transformed into finite difference form (see also Abbott (1979),
Cunge et al. (1986), Hirsch (1990) and Toro (1999)). Taking a -point as the only
feasible choice for the transformation of the continuity equation, the finite
difference form relates three successive unknowns uin-+1/1 2 , V in +1 and uin++1/1 2 defined at
the time level (n+1) t to known values at time level n t as

V in +1 - V in * hin+1/ 2 uin++1/q 2 - *hin-1/ 2 uin-+1/q 2


+ =0 (5.18)
Dt Dx
where
u n +q = (1 - q ) u n + q u n +1 (5.19)
and

t = the time step along the t-axis (s);


x = the space step along the x-axis (m);
n = a superscript denoting the time step number along t;
i = a subscript denoting the space step number along x;
= the time step weighting coefficient.

The symbol * in Equation (5.18) indicates that the value of h at this grid point has to
be approximated by its nearest available upstream value along the grid. For a
positive flow direction, for example, this means that hi+½ is approximated by the
value of hi.

Equation (5.18) can be reformulated as

a1i uin-+1/1 2 + b 1i V in +1 + g 1i uin++1/1 2 = d 1i (5.20)

where 1i, 1i, 1i and 1i are the coefficients of the linearized implicit finite
difference scheme.

Equation (5.18) can also be rewritten to provide a choice of time step that guarantees
the computation of positive water depths, as follows

æ Dt ö n n +q Dt
hin +1 = ç1 - uin++1/q 2 ÷ hi + ui -1/ 2 hin-1 (5.21)
è Dx ø Dx

For

uin-+1/q 2 ³ 0 and uin++1/q 2 ³ 0 (5.22)

Equation (5.21) shows that for these positive water velocities and for positive water
depths, the velocity Courant number condition

WL | Delft Hydraulics 89
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Dt Dx
uin++1/q 2 £ 1; or Dt £ (5.23)
Dx uin++1/q 2

can be applied to provide a sufficient condition for obtaining positive water depths at
the new time level. This implies that for the time step limitation of Equation (5.23)
newly computed water levels can never fall below the bottom of the channel. As a
consequence, no artificial bottom slots or other arrangements are required to avoid
numerical robustness problems for small water depths. A similar condition can be
derived for negative flow directions.

Following the same procedure as applied to the continuity equation, the momentum
Equation (5.17) can now be defined at a velocity point, by the finite difference form

uin++1/1 2 - uin+1/ 2 V in++1q - V in +q uin+1/ 2 uin++1/1 2


+ a (u , u ) + g
n n
+ cf =0 (5.24)
Dt Dx * n
hi +1/ 2

where the symbol * has the same meaning as in Equation (5.18) and a(un,un) is a
generalization of the discretization of the convective momentum term.

Appropriate formulations for a(un,un) follow from the physical conditions, as


detailed by Stelling and Duinmeijer (2003). In the paper it is shown that the correct
formulation of the convective momentum term depends on the way in which the
convective speed of momentum is interpolated on the grid. As a rule, the
discretization is based upon a transformation of the convective momentum term to

¶u 1 ì ¶ (uq) ¶q ü
u = í -u ý (5.25)
¶x h î ¶x ¶x þ

and its discretization

1 æ ui +1 q i +1 - ui q i ö
* *
¶u q -q
u ; ç - ui +1/ 2 i +1 i ÷ (5.26)
¶x hi +1/ 2 çè Dx Dx ÷
ø
where

hi + hi +1 q +q
hi +1/ 2 = ; q i = i -1/ 2 i +1/ 2 and qi +1/ 2 = *hi +1/ 2ui +1/ 2
2 2

Also here the value of *h has the same meaning as in Equation (5.18) The values of
*
ui are missing at -points on the grid and are approximated by first order unwinding
as

qi -1/ 2 + qi +1/ 2 qi -1/ 2 + qi +1/ 2


*
ui = ui -1/ 2 , if ³ 0 and *
ui = ui +1/ 2 , if < 0 (5.27)
2 2

For positive flow direction this yields the simple expression

WL | Delft Hydraulics 90
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

_
¶u q æu -u ö
u ; _ i ç i +1/ 2 i -1/ 2 ÷ (5.28)
¶x h è Dx ø
i +1/ 2

This momentum conservative approximation is applied in cases of gradually varied


flow or in flow expansions. Referring to Stelling and Duinmeijer (2003) again, this
formulation should not be used in strong flow contractions, as this will breed energy.
Instead, the energy head conservation formulation should be used, given by

¶u ¶V ¶ æ u2 ö
u +g =g ç +V ÷ (5.29)
¶x ¶x ¶x è 2 g ø

and, for positive values of u, the energy head conservative upwind discretization

¶u ui2+1/ 2 - ui2-1/ 2 1 æu -u ö
u ; = ( ui -1/ 2 + ui +1/ 2 ) ç i +1/ 2 i -1/ 2 ÷ (5.30)
¶x 2Dx 2 è Dx ø

For negative flow velocity, this complete expression is shifted one grid point in the
positive x-direction. Comparison of Equations (5.28) and (5.30) leads to the
conclusion that the difference in obtaining momentum or energy conservation lies in
the way in which the convective velocity of momentum is interpolated from the flow
field.

Terms in Equation (5.24) including (5.26) or (5.30) can be collected to give the
generalized relation

a 2i V in +1 + b 2i uin++1/1 2 + g 2i V in++11 = d 2i (5.31)

By successive elimination of all equations at u-points, the remaining set is given by

a i V in-+11 + bi V in +1 + g i V in++11 = d i (5.32)

and is solved by applying elimination or conjugate gradient techniques, as discussed


further down in this contribution.

The numerical discretization is second order accurate for all terms of the de Saint
Venant equations, except for the convective momentum term, which is first order
accurate. For many practical applications the convective momentum term is of lesser
importance and this lower discretization accuracy is quite acceptable then. However,
nearly second order accuracy can be achieved by up winded second order
extrapolation of u-values, combined with slope limiters, as described by Stelling and
Duinmeijer (2003), again.

In rapidly varied flow the convective momentum term becomes locally dominant.
Modelling these types of flow requires an appropriate implementation of the
convective momentum term. By doing so, Delft Hydraulics’SOBEK package has

WL | Delft Hydraulics 91
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

been enabled the highly accurate and robust modelling of phenomena such as
supercritical flow in steep channels and moving hydraulic jumps.

The numerical scheme given by Equations (5.20) and (5.31) only provides linear
equations at internal grid points. At channel boundaries, additional equations are
required. As the de Saint Venant equations are of second order hyperbolic type, one
boundary condition is required for each of its characteristic lines entering the
computational domain. Boundary condition requirements for 1D river models were
discussed extensively by Cunge, Holly and Verwey (1986). Here we will limit
ourselves by stating that, in most practical applications, inflowing discharges are
specified at the upstream ends of channels entering the flood model domain and
water levels or rating curves at channels leaving the model domain. At internal
boundaries, such as channel junctions, usually a modified continuity equation is
applied, jointly with water level compatibility at all channel boundaries at that
junction.

5.5 Description of hydraulic structures


Nearly all flood simulation models are dealing with hydraulic structures, such as
dams, weirs, bridges etc. The length of the hydraulic structure along the x-axis is
usually supposed to be negligible on the scale of the river or channel and therefore
the structure can be seen as one single point along x, where both water- and energy
levels are discontinuous and the de Saint Venant equations do not apply. At this
point a relation is established between the upstream water level, the discharge
through the structure and the downstream water level. In this relation it is assumed
that the upstream water level is taken at the nearest point just upstream of the
structure, where vertical accelerations can still be neglected. Similarly, the
downstream water level is taken at the nearest location just downstream of the
structure, where the flow can be seen as nearly horizontal and the structure flow no
longer contributes to the energy dissipation. Usually, it is also assumed that storage
of water around the structure is negligible, so that the relationship is applicable at
any moment in time. On this basis, there is a variety of ways in which the flow
through hydraulic structures can be formulated:

· in most cases hydraulic structure descriptions are based upon empirical laws
relating the discharge through the structure to the upstream and downstream
water level. The relation has the general form

Qcrest = Q (V up , V down ) (5.33)

where up is the water level upstream of the structure (m), Qcrest the discharge
through the structure (m3/s) and down the water level downstream of the
structure (m). This formulation includes the possibility of using energy levels
instead of water levels. Equation (5.33) may be linearized to

a V up + b Qcrest + g V down = d (5.34)

WL | Delft Hydraulics 92
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where , , and are the local values of non-linear coefficients at a given state
of the flow. A wide variety of structure descriptions is available from literature,
e.g. the classical books of Chow (1959), Henderson (1966) and, more recently,
Chanson (1999);
· as an alternative, the state of the flow in the section just upstream of the
structure, where the flow is contracted and vertical accelerations occur, can be
described by an energy conservation principle. Similarly, the state of the flow
just downstream of the structure, where the flow expands with energy losses
associated to it, can be described by a momentum and impulse balance
relationship. By internal elimination of unknowns in these relationships, the
water level at the structure crest or at its most contracted section can be
eliminated and a relation similar to Equation (5.34) is obtained;
· in cases where it is difficult to define the state of the flow in terms of equations,
laboratory experiments may be set up to define a matrix relating upstream and
downstream water levels to the structure discharge. As an alternative, the
structure may be modelled in detail by a 3D numerical code, leading to a similar
set of matrix coefficients. Conditions are that a fine grid is used, and that the
code is based upon an appropriate numerical description of convective
momentum terms and the effect of turbulence.

In all cases shown above, the structure equations could be based upon energy levels,
instead of water levels. Through suitable transformations, these relations can be
reworked to the form of Equation (5.34) Similar descriptions can be made for local
energy loss descriptions along a channel.

Appropriate linearization of Equation (5.33) or of the other structure flow


descriptions, leads to a numerical scheme of the form of Equation (5.31). In the total
set of equations for channel flow this structure relationship replaces the momentum
equation that is generally applied between successive -points.

In the case of a closed structure or a discharge specified by a pump, the coefficients


and of Equation (5.31) are both zero and the structure presents itself as an
internal boundary condition with a discharge given. In the case of free flow in the
positive x-direction, only the coefficient equals zero, showing also in the
numerical relationship that the structure discharge is only dependent on the upstream
water level.

In literature an abundant number of structure equations can be found (Chow, 1959,


Henderson, 1966 and, more recently, Chanson, 1999).
Implementing these in a numerical code requires attention for discharge
compatibility between the various flow states, especially at the transition between
free and submerged flow. Incompatibility in the definition of the discharge may lead
to oscillations in the computed flow, in particular at the moment of flow reversal.
Problems may be suppressed by defining some inertia to the structure flow. This is
implemented by adding a small term to the coefficients and . The numerical
structure behaviour is generally better when applying an energy relationship in the
accelerating flow section upstream of the structure combined with a momentum-
impulse relationship for the section downstream of the structure.

WL | Delft Hydraulics 93
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Where appropriate, energy losses may be added to the description of flow in the
contracted section. Internal elimination of local variables leads, again, to a
relationship as given by Equation (5.31) . In all cases, the different flow states only
affect the computation of the coefficients , , and and have no bearing on the
solution algorithm of the resulting system of equations.

5.6 Topological model schematisation


Mathematical models for the simulation of flow in river- and channel networks are
based upon the De Saint Venant equations, hydraulic structure equations and initial-
and boundary conditions. The modeller has to simplify the complete system to a
choice of schematization objects which will give a satisfactory representation of the
physical behaviour of the hydraulic system. Two processes are of primary
importance: storage and conveyance.

The actual model construction will start with the selection of the network branches:
the topological schematisation. In this process it will be decided up to which level of
detail network branches are to be included in the schematization. If too many minor
channels are included, the amount of work in data preparation may be unnecessary
high. In addition, model simulations may become too slow. The practice of good
model construction requires a good insight into the relative importance of these
contributions.

Starting with conveyance, let us discuss the following issues:

· what is the relative contribution of a channel to the overall conveyance of the


system? For a good judgement, the Manning equation shows that if two channels
give flow into the same direction, their relative importance is found by
comparing their conveyance. Applying the Manning concept, Equation (5.11)
shows, for example, that for similar roughness, a channel with half the width and
depth of a reference channel, has only 16 % of the discharge capacity of that
reference channel;
· can a minor channel form a short cut? Not only the conveyance of a channel is
important. Certain channels may provide a short cut to the flow from one part of
the network to the other. Even for small conveyance values the discharges may
become substantial due to a steep water level slope. Special attention has to be
given to a possible over bank flow at higher stages;
· is the water level always the same all over the channel cross-sections? If this is
not the case it may be decided to split the channel up in two or more parallel
branches and describe cross flow between the parallel channels via cross-
channels or hydraulic structure descriptions. In general, such schematisations
lead to significantly more complex model networks.

This last point is also important for the correct representation of storage in the
model. Natural levees along rivers may delay the storage in the flood plain and
hence, affect the celerity of flood wave propagation along the river and the
dampening of flood wave peaks.

WL | Delft Hydraulics 94
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Other questions addressed are:

· is a storage area well connected to the network? This means, for example, that it
can be fed with the correct conveyances by flow from all relevant directions;
· are there any barriers in the flood plain? If part of the flood plain is schematized
as a storage element, it may contain various embankments or levees. Examples
are levees raised around paddy fields. During floods, part of the described
storage area may not get flooded at all.

Finally, lateral flows may have their impact on the topological schematisation as
follows:

· should we extend the schematisation to include part of a tributary? For example,


how far does the backwater influence the tributary and does this effect justify the
inclusion of part of the tributary in the topological schematisation? If a
measuring station is available to provide the lateral flow time series, it may be
quite far from the main river. In this case the schematisation may be extended to
include this location;
· should we couple other models to our hydrodynamic model? For example, the
exchange of surface water with the sub soil may be substantial or it may at least
be important in relation to the objective of the model development. Rainfall-
runoff models may become an integrated part of the total description. In flood
wave propagation the role of infiltration of surface water into the sub soil may
have to be included in the model.

5.7 Hydraulic model schematisation


Once decisions have been made about the channel network schematisation and its
various model elements, choices have to be made about equations and their
associated data. In most cases, channel flow will be based upon the description of
the full De Saint Venant equations, although also mixed models may be constructed.
It could make sense, for example, to describe the flow in upstream, steep river
branches with hydrologic models or artificial neural networks (ANNs), whereas in
the principal rivers the full De Saint Venant equations are applied. In all cases,
however, the models must include representative values for channel storage and
conveyance, whether these are specified directly or hidden in the parameters of a
hydrologic model. In principle, the model must compensate for conveyance or
storage which has been neglected in the topological schematisation.

One important decision is the choice of distance step x, which should be based
upon a good representation of the hydraulic processes. The choice is based upon:

· The wave length modelled. A rule of thumb is to model waves with at least 50 to
100 grid point along important wave components;
· The representation of the local variations in hydraulic parameters, such as
sudden contractions and expansions;
· Placement of grid points closely around hydraulic structures or other
discontinuites in the system.

WL | Delft Hydraulics 95
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Channel conveyance is described at cross-sections. Under all circumstances these


cross-sections must be lines perpendicular to the direction of mean velocities. In the
case of meandering rivers these cross-section alignments have to be defined on the
basis of sound engineering judgement. A complicating factor is the change of
velocity vector directions with changing stage.

For man-made channels the specification of cross-section parameters is rather


straight forward. As the dimensions are usually quite accurately known, the most
sensitive input parameter is the roughness coefficient. For simple channel
geometries, the use of the hydraulic radius as an approximation of a representative
depth is quite acceptable. The hydraulic radius is defined as the surface of the
conveying cross-section divided by the wet perimeter (Equation (5.11) or (5.12)) and
assumes a uniform distribution of the shear force along this wet perimeter. For
cross-sections with varying depths this assumption is incorrect. The local shear force
in the cross-section is a linear function of the local depth. For this reason, the
conveyance of compound channels has to be based upon a summation of the
conveyances of individual sub-sections where for each of them this shear force is
more or less uniform (Equation (5.9)).

An advantage of such integration is that each sub section can be given its own
roughness value. Another advantage is that by keeping track of the individual
contribution of the sub sections to the overall conveyance of the channel, computed
discharges can be redistributed across the section to provide water velocities for
each sub section. This ability may be important in simulating water quality or
morphological processes in flood plains where the use of local velocities is required.
In the integration, contributions of sub sections with small conveyance are without
any loss of accuracy added to those of the other sub sections. There is no need to
neglect the conveyance of shallow and highly resistant parts of the cross-sections as
these may still give substantial contributions during floods or may be important in
the overall assessment of the morphologic or water quality behaviour.

One of the basic assumptions in a one-dimensional schematisation is the use of a


constant water level slope all along the cross-section. In meandering channels,
however, the slope may be quite different for different sub sections. One way to
include this effect in the derivation of the conveyance is by giving each sub section a
weight in the integration, as a function of the distance to its equivalent part in the
next cross-section. Equation (5.11) will then be modified as follows

( DV ) ( DV )
½ jj ½
1
Q = K = ån Aj R j 2/3

( Dx ) ( Dx )
½ ½
j =1 j
j

or
jj
Aj R j 2 / 3
K = å (5.35)
n j ( Dx j / Dx )
½
j =1

WL | Delft Hydraulics 96
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

where x is the distance for which the cross-section parameters are representative
along the x-axis and xj a similar value for sub section j.

Storage parameters usually are given as a function of stage and linked to cross-
sections. For uniform channels, as often designed in irrigation and drainage systems,
the storage width is equivalent to the flow width. Storage width data are simply
extracted from cross-section information. In natural rivers, however, meandering and
irregular flood plain topography requires more complex procedures. § 5.2 gives an
extensive discussion of the correct definition of the storage width parameter bs. It
follows that storage parameters have to be extracted from information provided by
topographic maps, currently mostly available in the form of digital elevation models
(DEMs) in GIS. For successive compartments along the river axis, approximate water
level slopes have to be assumed to extract from these maps storage area as a function of
stage. Division of these areas by the length of the compartment along the x-axis,
provides the parameter values for bs. For models of lesser economic value, procedures
may be simplified, if needed.

Storage connected to channels which have been neglected in the topological


schematisation must be added to the model as additional storage. In the cross-sections
the compensation can be made in the form of additional storage width. It can also be
introduced in the schematisation in the form of additional storage areas at nodes. It is
recommended to keep track of the changes that this correction phase has given in
parameter values derived during the primary phase of schematisation. It is always
advised to keep good records of all the steps taken in the development of a model and
the way the model parameters have been defined. Without such records it is practically
impossible to introduce future improvements or modifications correctly. This is even
more so when various persons are involved in the model development and/or the
development takes place over a considerable length of time.

Hydraulic structures may be described by their empirical relationships, by the


application of an energy conservation principle upstream and a momentum
conservation principle downstream or by specifying the discharge water level
relationships in matrix form. In case of empirical relationships, care must be taken that
the parameters describing free flow and submerged flow, successively, give a
consistent computed discharge at the moment of transition from one flow state to the
other. If applicable, additional entrance energy losses must be specified. Composite
cross-sections of the structure require an adaptation of the topological schematisation
by defining parallel structures. Similar to the definition of conveyance, the total
discharge through the composite structure is defined by adding up the contributions of
the individual structure components.

5.8 Boundary- and initial conditions


In relation to boundary conditions two issues are important: the location of a model
boundary and the condition applied at this point. The location depends on the model
objective. For models meant to study hydraulic problems, a strict rule for a boundary
location is that the boundary data given at this point should not affect the outcome
computations of various scenarios.

WL | Delft Hydraulics 97
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

In other words, the model changes introduced with these scenarios and the resulting
changes of the hydraulic state at the boundaries of the models should not reflect
back on the study area. Exceptions to this rule are rare, but may be justified in some
cases of water quality modelling, for example.

There are various ways in which this can be achieved. Water levels at a sea
boundary of a river model are generally not affected by the river discharge. Usually,
it is safe to use a sea boundary as a boundary location in a river model and provide
water levels as boundary condition.

In other cases it must be guaranteed that the downstream river boundary is located at
a sufficient distance from the area studied. How far, depends on how an error, or a
change in the downstream boundary condition, affects the area under study. In the
first place, this depends on the magnitude of the error or change. In the second place
this depends on the hydraulic characteristics of the river. A good measure for the
required distance is the length of a backwater effect. In case of doubt it is
recommended to do a sensitivity analysis on the effect of an error in the downstream
boundary condition.

At upstream boundaries the same reasoning applies. In this case the scenario applied
should not reflect back on the data or condition applied at the upstream boundary.
Also here, the required distance follows from an analysis of backwater effects.

The type of boundary condition follows from the theory of characteristics, stating
that one boundary condition has to be given for each characteristic entering the
computational domain. In practice, however, this always means that one boundary
condition is given at each boundary of the model. For sub critical flow this is usually
a discharge hydrograph at an upstream end and water levels or a rating curve at a
downstream end.

A rating curve at the upstream end will usually lead to model instabilities as such
condition will be nearly equivalent to the information contained in the channel
topography. In this way, this boundary condition does not provide additional
information that is not already available and the state of the flow remains
undetermined. A rating curve conflicting with the topographic information leads
mathematically to indeterminacy of the solution and hence, to instabilities.

At downstream boundaries, usually water levels or rating curves are given. In river
applications one could define this as a weak boundary condition, as it will be shown
in Chapter 6 that the wave propagating in a river is usually an advection type
phenomenon, only requiring an upstream boundary condition. This implies that for
the use of the full hydrodynamic equations the downstream boundary condition
becomes redundant.

For similar reasons, it is sufficient in practice to give only one upstream boundary
condition for super critical flow. The other condition is already implied in the
relation between topographic data and the water level slopes following from this.

WL | Delft Hydraulics 98
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

At the downstream boundary, any condition given, will at worst lead to incorrect
results over a small stretch of the downstream region. As a rule, the downstream
boundary condition given will turn the outflow locally into a state of sub critical
flow.
Initial data are simplest supplied by starting up flow simulations on a dry bed or a
low initial water depth. In the case of tidal models this specification is replaced by a
constant mean sea level initially and zero discharges throughout. In tidal flow,
usually two tidal cycles are sufficient to arrive at a consistent set of initial data, at
least in terms of water levels. Most modelling systems provide the option of
generating a set of consistent initial data and writing this state to a file to be used as
a so-called hot-start file for further simulations.

5.9 Model calibration and validation


Calibration of a model is the process of removing, or at least reducing, the
uncertainties in the choice of model parameter values. In principle, nearly all model
data can be collected from the field, except for the roughness values. These
parameter values can only be obtained indirectly by comparison of measured and
computed variables, such as water levels and discharges. In principle, a model
calibration could focus only on the optimization of roughness coefficient values.

However, in practice there are uncertainties also in relation to cross-section data,


flood plain topography, lateral flow data and inflowing hydrographs. This means
that in a calibration also the correctness of storage parameters and measured
hydrographs will have to be checked and possibly corrected. Part of the uncertainty
is also caused by the schematization of the real system into the model. Neglected
storage and conveyance and an inappropriate compensation for these missing effects
also influence the calibration results.

It is possible, then, that corrections are applied to the wrong parameters and still lead
to reasonable calibration results. As will be shown in Chapter 6, flood wave
propagation celerity and dampening both depend on storage and conveyance. An
error in the storage could, therefore, be corrected through the channel resistance and
still improve calibration results. It is, however, not certain that the same
improvement is maintained under extrapolated conditions.

For river channels, calibration of roughness data should proceed from lower
discharges to higher discharges. Only after the roughness values in the main channel
have been calibrated, those of the flood plain should be determined.

Model validation is meant to provide confidence in the quality of a calibration. In


principle, model validation should be applied on a case representing an extrapolation
of the calibration events. Many models are developed for use beyond the range of
data for which they could be calibrated. Flood models, for example, often are used
for studies on flood frequencies of several hundreds of years and it cannot be
expected that data of such events is available for model calibration.

Moreover, one of the principal problems with model calibration is that results that
have been measured during low frequency events, are not very reliable.

WL | Delft Hydraulics 99
Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Rating curves for these low frequencies are rather unreliable as measurement
conditions are difficult and the staff of catchment authorities generally have other
worries than getting the most accurate discharge measurements during an extreme
storm event.

Only water levels have a reasonable chance of being reliable, also during extreme
conditions. A possibility, therefore, is the extrapolation of rating curves on the basis
of the use of local 2D models, based upon an accurately measured topography and
calibration of roughness values during the higher frequency events.

However, this approach is not common practice yet. On the other hand, Public
Works in The Netherlands uses detailed 2D models of the flood plains of the
principal rivers for their calibration of 1D models. These 1D models can than be
used with more confidence in the range of extreme events. It should be realised that
this procedure is feasible only if these 2D models are based upon very small size
grid cells.

WL | Delft Hydraulics 100


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

6 Mathematical modelling of floods

6.1 Introduction
All over history, floods have challenged scientists and engineers to master this
phenomenon, both by developing mathematical descritions and by taking flood
protection measures. Over the past decades, these efforts even have increased. In the
first place, because a rapidly increasing world population has shown a strong drive
to settle in flood prone coastal zones and river flood plains. In addition, there is an
increasing awareness about climate changes which appear to bring more and more
precipitation to river catchments. There also are numerous intellectual challenges, as
scientific and technological developments in a wide range of fields open up many
new ways of supporting flood related studies. In particular, important progress has
been made in computer speed and in new measuring technologies.

Although numerical methods for solving unsteady flow equations were developed
already half a century ago, e.g. Arakawa (1966), it was only recently that numerical
techniques behind flood simulation models have reached an acceptable level of
perfection. Hydraulic engineers and hydrologists are now able to model flow in
channels and over flood plains, irrespective of their bathymetric or topographic
complexity; irrespective of the number and location of embankments for flood
protection, roads and railways and irrespective of the number and complexity of
hydraulic structures and the way we control these.

6.2 Flood model requirements


Flood simulation models may have different requirements, depending on their
objective. Criteria for the selection of the appropriate tool are often based on:
engineering staff time needed for model development, overall consultancy time for
product delivery, speed of computation, completion time for a simulation, accuracy
level of results, data requirements, numerical robustness, user-friendliness of the
software and possibly others, depending on the objective of the model. These
objectives may be related to flood risk analysis, flood forecasting, flood control and
be based upon a variety of causes, such as storms, dam or dike breaks, hurricanes,
typhoons or similar low atmospheric pressure phenomena. Recently (2004) attention
has also been drawn once again to the devastating effects of tsunamis. All these
application areas of numerical models have their own requirements, as will be
discussed briefly in the sequel.

In many countries, insurance companies are using flood risk maps, sometimes based
upon relatively simple and quick estimates obtained via simple rules in GIS. In these
cases it is assumed that the value of insured property does not justify the more
complex laws defining the detailed flow of water.

WL | Delft Hydraulics 101


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

If the economic interest is greater, a first improvement is found by applying 1D


(one-dimensional) steady flow models with GIS post processing to develop
topography based flood frequency contour lines (e.g. FEMA procedures:
www.floodmaps.fema.gov/fhm/). However, there is a tendency now to base flood
risk analysis on more detailed combined 1D and 2D unsteady flow models for flood
prone areas with valuable assets and a complex infrastructure. Federal and local
governments have also become more aware of the potential of using such models for
evacuation planning. In The Netherlands, for example, more than 60 % of the
country is subject to flood risk and for most of these areas integrated 1D and 2D
hydrodynamic models have been developed to study the effects of potential dike
breaks and to provide guide lines to authorities in setting up evacuation plans.
Besides producing flood depths, these models have to be capable of providing
accurate estimates of flood wave propagation celerities over dry beds.

Flood forecasting sets quite different requirements. Speed of producing a forecast is


one of the most important criteria, especially in areas where flash floods occur. For
this reason, numerical models behind a river catchment flood forecasting system are
usually 1D hydrodynamic models, gradually replacing the simpler hydrological
routing techniques. There is a tendency to include partly 2D hydrodynamic models,
which is already common practice in flood forecasting systems for coastal areas and
seas. Numerical models for flood forecasting are usually embedded in a flood
forecasting platform, such as the Delft FEWS system (Werner et al., 2004), which
has recently been installed in the UK to provide flood forecasts for nearly all river
basins in the country.

Important criteria for numerical models supporting flood control are accuracy,
flexible schematization options, numerical robustness and consultancy time for
model development and use. Currently, state-of-the-art for flood control is the use of
combined 1D and 2D models (e.g. Hesselink, 2003). The former use of flood cells
has been replaced by complete 2D flow descriptions, whereas sub-grid channel flow
is still better described in 1D. Flood control models should be based upon reliable
physical descriptions and schematizations, as part of their use is in extrapolation of
calibrated models to extreme situations which have never occurred. One of the
reasons to build models for flood control is the study of downstream impacts,
especially cross border effects. Downstream impacts of flood control are changed
flood wave celerity and changed flood peak attenuation. Higher flood wave
celerities result from deepening of the river and the construction of embankments.
This, in turn, leads to increased peak floods downstream. The construction of flood
retention areas has opposite impacts and may be used to compensate the negative
impacts. Model selection criteria then follow from the detail in which potential
economic, environmental and social impacts have to be studied.

The analysis of floods caused by dam- and dike breaks requires extremely robust
numerical methods, especially for the description of flooding of dry areas and the
correct propagation of the wave front. Moreover, model accuracy, partly based upon
the ability to describe the full hydrodynamic equations, is important, as will be
discussed in the section on software and model validation. As dam- and dike break
simulations are nearly always made for the prediction of their potential effects, data
for model calibration is rarely available.

WL | Delft Hydraulics 102


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The quality of the model fully depends on its descriptive capabilities of the physical
system in terms of topographic and roughness data, the representativeness of the
equations and the numerical methods applied. However, it has to be kept in mind
that the overall model accuracy also follows from the quality of the description of
the dam failure mechanism and the assumptions made here.

Floods generated by the passage of low atmospheric pressure zones such as


hurricanes, typhoons and the geologically induced tsunamis require the modelling of
2D flow in coastal zones, seas and oceans and may set requirements such as the
description of Coriolis forces, the use of spherical coordinates and curvilinear grids,
the specification of moving atmospheric pressure fields, special ways of handling
initial data etc. A possible integrated use of 1D, 2D and 3D models may provide
advantages here.

6.3 The role of new data collection technologies


Models need good quality data if one wants to draw reliable conclusion from their
use. In addition to developments in numerical methods, new technologies for the
collection of data have recently led to complete changes in the selection of the type
of numerical models. In particular, the development of GPS and DGPS technology
has led to far cheaper methods of collecting bathymetric and topographic data
(Moglen and Maidment, hsa025). In turn, this has led to the gradual replacement of
the hydrograph based hydrologic models by the bathymetric and topographic based
hydrodynamic or hydraulic models. Similarly, the collection of detailed digital
terrain data in river and coastal flood plains has led to the replacement of 1D models
by 2D models.

Let us first consider the impact of LIDAR. The use of this laser technology,
scanning the earth surface with laser beams from airplanes or helicopters, has
provided the means to generate highly accurate digital elevation models at relative
low cost. As an example, the whole area of The Netherlands has been remapped
over the past years, with an accuracy of approximately 10 cm in the vertical at a
density of 1 point per 16 m2. Total cost of this project was approximately 10 million
euro, or approximately 250 euro per km2 (Verwey, 2001). This new topographic
information has been essential for the flood risk analyses and the evacuation
planning studies mentioned earlier.

Also river bathymetries are obtained at relative low cost now by boats equipped with
multibeam echo sounders. Also here, the position of the boat is recorded via DPGS,
while the spatial sound signals record bottom depths relative to the boat at an
accuracy of approximately 10 cm in the vertical. At a typical boat speed of 4 m/s and
a swat width of the order of magnitude of the river depth, the bathymetry of quite
large river beds can be obtained in a few days. Experiments are also being made
with laser beams in the green range, as these are able to pass through clean water
and enable the application of LIDAR technology also for river bathymetries.

In similar ways other technologies are advancing rapidly enabling large amounts of
data to be collected at relatively low cost.

WL | Delft Hydraulics 103


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Worth mentioning are the ADCP (Acoustic Doppler Current Profiler) technology for
tidal discharge measurements and the increased precision of spatially distributed
precipitation measurements with radar.

6.4 The nature of flood wave propagation


The set of Equations (5.1) and (5.11) forms the so-called kinematic wave
approximation for flood propagation, if the slope I is taken as the bed slope. After
substitution of Equation (5.11), or the kinematic wave part of Equation (5.15), into
Equation (5.1) and neglecting the lateral flow term, a further simplication is
achieved of the form

¶Q ¶Q
+c =0 (6.1)
¶t ¶x

with a flood wave celerity c (m/s) expressed as

1 dQ
c= (6.2)
bs dh

It is important to realise that, as an essential assumption in the derivation of this


kinematic wave form, the channel bed slope has been used in the conveyance
relationship. Equation (6.1) has the form of an advection equation which expresses
that flood waves propagate with a celerity c which is inversely proportional to the
available channel storage width bs (m) and a linear function of the derivative of the
local flow rating curve. The characteristic celerity of this kinematic wave is lower
than the celerity of the dynamic wave characteristic in the same direction. As
discussed by Abbott (1979) it is this mechanism that leads to roll waves at flood
wave fronts, limiting their propagation speed (see also Stoker, 1957).

The first order partial differential equation (6.1) also expresses that along its
characteristic celerity c the discharge remains constant, and so does the peak of the
flood wave. In other words: there is no dampening effect of the flood peak. Though
this is approximately true for rivers with steep slopes, the stretches with milder
slopes require a lesser simplification of the De Saint Venant equations. Following,
for example, Chaudhry (1993) and defining I of Equation (5.11) as the water level
slope z, substitution of the last three terms of Equation (5.15) into Equation (5.1)
gives the so-called diffusive wave approximation

¶Q ¶Q ¶ 2Q
+c =D 2 (6.3)
¶t ¶x ¶x

with a flood wave diffusion coefficient D (m2/s) derived as

K
D= (6.4)
2bs I

WL | Delft Hydraulics 104


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Including the lateral flow term the diffusive wave approximation reads

¶Q ¶Q ¶ 2Q
+c = D 2 + cq (6.5)
¶t ¶x ¶x

Returning to Equation (5.15) again, the variable I represents the bed slope in the
case of the kinematic wave approximation and the water level slope in the case of
the diffusive wave approximation. The description based upon the full set of
Equations (5.1) and (5.15) is defined as the full dynamic wave description.

6.5 Deformation of flood waves

6.5.1 The role of varying celerities

Returning to Equation (6.2), it is readily seen that for constant celerity at all water
levels, there is no deformation of the flood waves. The waves are just translated
along the river axis. However, in practice the wave celerity c generally increases
with increasing river discharge, especially when the river banks are steep (Figure
6.1a). Moreover, when the flood waves arrive at bank level, the storage width often
increases more rapidly than the increase in dQ/dh, resulting in a temporary dip in
this celerity function (Figure 6.1b). When the water level rises further, the celerities
start increasing again.

Figure 6.1 Flood wave celerity for various cross-section shapes

The higher celerities for increasing discharges cause a steepening of the wave at the
front and a stretching of the falling limb of the flood hydrograph (Figure 6.2). This is
the most common form of flood wave deformation in natural river systems.

Figure 6.2 Flood wave deformations resulting from celerity varying with stage

WL | Delft Hydraulics 105


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

6.5.2 The role of the diffusion term

Equation (6.3) shows that along the characteristic line c=dx/dt the flood peaks are
dampened by the integral of the diffusion term. As the diffusive wave approximation
of the De Saint Venant equations may include other effects as well (e.g. the form of
Equation (6.5)), extending the number of terms of the right hand side, we will
discuss these various contributions in a fractioned step approach.

Returning now to the right hand side terms of Equation (6.3), the influence of the
diffusion fraction can be written as

2
dQ ¶Q
( )1 = D 2 (6.6)
dt ¶x

which gives the equation for flood peak attenuation as

æ ¶ 2Q ö
dQ1 = D ç 2 ÷ dt (6.7)
è ¶ x øpeak

The diffusion mechanism is commonly explained by the fact that the water level
slope during rising water level is steeper than the bed slope. Consequently, the
discharge at the wave front is higher than might be expected from the rating curve
and the water balance in that section forces a lowering of the wave peak. This is
further enforced by the lower discharges at the rear of the wave as compared to the
discharge which would follow from the stage-discharge relationship. Here, the water
level slopes are smaller than the bed slope (Figure 6.3).

Figure 6.3 Illustration of the effect of varying water level slopes on flood wave deformation

Returning to Equation (6.7), it is readily seen that this fraction gives a dampening of
the wave peak, as the term ¶2Q/¶x2 is negative, while it leads to an increase of the
discharge at the front and at the rear of the wave, where ¶2Q/¶x2 is positive.

Observation of the diffusion coefficient given by Equation (6.4) leads to an


interesting paradox. It is a well known phenomenon that an increase in storage also
increases the dampening of flood waves. This, however, does not follow
immediately from Equation (6.4), which shows a decrease in the wave diffusion
coefficient with increasing storage width of the river. It should be realised that an
increase in storage width also decreases the flood wave celerity, leading to a
proportional decrease in the flood wave length. Consequently, the contribution of the

WL | Delft Hydraulics 106


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

term ¶2Q/¶x2 shows a quadratic increase in the wave dampening. Furthermore, the
decrease in the wave celerity increases the travel time, or the time over which
Equation (6.7) is integrated, over a given stretch of the river, compensating for the
decrease in the diffusion coefficient. Taking all these effects into account, an
increase in the storage width by a factor two, gives a total increase in wave
attenuation of a factor four.

6.5.3 The role of the lateral flow terms

Including also the effect of lateral flow, Equation (6.5) shows its influence.
Applying, again, the fractioned step approach, the contribution of the lateral flow to
the change in peak discharge is given by

dQ
( ) =cq (6.8)
dt 2
or

dQ 2 = cq peak dt = q peak dx (6.9)

Equation (6.9) simply tells us that the peak value of the flood wave is increased by
the total lateral flow added to the flood as its peak travels along the river.

Interesting conclusions can now be drawn on the case where flood water infiltrates
into the soil as the flood passes the flood plain. As the infiltration is a loss of water,
the lateral flow is negative and leads to a reduction of discharges along the paths
given by the characteristics. A typical infiltration function along the flood wave
would show the highest water losses during the passage of the front of the wave,
which shows up as a pronounced effect when the bed material is highly permeable.
A high porosity of the bed material, gravel for example, causes a rapid increase in
the infiltration when water is made available by the wave. During the further rise of
the water level, the stream width increases and increases the storage volume
available for infiltration. At the passage of the peak, the infiltration is more or less
completed and the function drops down to zero.

A typical response of this infiltration is a pronounced steepening up of the flood


wave front. This is why such floods may have a very short lead time for warning the
population and may be quite destructive. A typical example is given by floods in
wadis. However, also in more moderate climates this effect is known. Note that it is
necessary in such cases to include in a model the effect of exchange with the
ground-water, as the typical wave deformation found in this case cannot be
calibrated by modifying friction parameter values.

6.6 Link to hydrologic flood routing models


Equations (6.1) and (6.3) are rarely used in discretized form anymore. However,
they are useful for the flood modeler as to provide insight into the physical nature of
flood wave propagation. Moreover, they link hydrologic and hydraulic flood routing

WL | Delft Hydraulics 107


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

techniques. By applying a Taylor’s series expansion to the Muskingum equation,


Cunge (1969) has demonstrated that the well-known hydrological Muskingum flood
routing technique emulates the solution of the advection-diffusion Equation (6.3).
This insight provides guidelines for a suitable choice of the Muskingum parameters
on the basis of the expressions for c and D (Equations (6.2) and (6.4), respectively).

The numerical Muskingum method belongs to the class of hydrological routing


techniques. These methods are based upon the notion that flood wave propagation
characteristics can be derived from measured flood hydrographs along the river,
rather than from detailed topographic information, as discussed in the WMO report
on flood forecasting models by Serban et al. (2005). The basis of these methods is
that flood wave propagation and diffusion behavior can be represented by a limited
number of parameters, which can be calibrated from observed hydrographs. Usually
there are only two parameters, as in the case of the Muskingum routing method. The
limitation of this assumption is clearly shown by analyzing the stage dependent
expressions for celerity and diffusion given by Equations (6.2) and (6.4). Although
the current methods for flood routing are far more precise, there are still many
situations where the use of hydrological flood routing techniques is still justified. In
many river catchments, and especially in tributaries, the collection of detailed
information on river bathymetry and flood plain topography is economically not
always justified, despite the emergence of relatively cheap new technologies.

An alternative to hydrological forecasting methods is provided by newly developed


artificial neural network (ANN) concepts (Minns & Hall, hsa018). Also this
technology relies on measured hydrographs to derive relationships between
inflowing and out flowing hydrographs, though in practice ANN’s are mostly
applied to the development of relationships between rainfall and river catchment
runoff.

Both hydrological routing and ANN techniques provide limited reliability for the
range of events outside those used for calibration. However, this is exactly the range
one is interested in when dealing with extreme flood events, which rarely occur and
for which measurements are even more rarely available. So even though ANNs and
other data-driven modelling techniques may prove quite valuable in e.g. determining
rainfall-runoff relations (Solomatine, hsa021), physically based descriptions, such as
provided by hydraulic routing techniques and based upon the full use of Equations
(5.1) and (5.2), offer better extrapolation possibilities than hydrological routing
methods or ANN based models. For this reason, we will focus on hydraulic
modelling techniques in the sequel. However, it should be realized that simplified
equations, such as Equation (6.5), remain useful, as these offer us a good insight into
the physical nature of flood propagation, in particular the concepts of flood peak
arrival time and the attenuation of peak discharges.

6.7 Two-dimensional modelling of floods


In the modelling of floods, flows often take short cuts through flood plains where
the 1D description may become quite inaccurate. This is even more the case for
dam- or embankment failures, where the flow may leave the flood plain completely
and inundate natural terrains. For this reason the two-dimensional (2D) shallow

WL | Delft Hydraulics 108


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

water equations will be introduced. Following the same principles as for 1D flow,
these equations read

¶h ¶ (uh) ¶ (vh)
+ + =0 (6.10)
¶t ¶x ¶y
¶u ¶u ¶u ¶ (h + zb ) u u 2 + v2
+u +v +g + cf =0 (6.11)
¶t ¶x ¶y ¶x h
¶v ¶v ¶v ¶ ( h + zb ) v u 2 + v2
+u +v +g + cf =0 (6.12)
¶t ¶x ¶y ¶y h

where we now also introduce the y-axis, orthogonal to the x-axis, with its flow
velocity v (m/s) associated to it. The friction term, with the dimensionless friction
coefficient cf, has in both momentum equations a shear force component derived
from the quadratic head loss description along a stream line of the 2D flow. Basic
assumptions are similar to those given for the 1D equations, as far as applicable in
this form of schematization.

Figure 6.4 Staggered grid for 2D flow simulations

Referring to the 2D grid shown in Figure 2, a volume conservative finite difference


form of the continuity equation is given by

V in, +j 1 - V in, j * hin+1/ 2, j uin++1/q 2, j - *hin-1/ 2, j uin-+1/q 2, j


+ +
Dt Dx
(6.13)
hi , j +1/ 2 vin, +j +q1/ 2 - *hin, j -1/ 2 vin, +j -q1/ 2
* n

=0
Dy

WL | Delft Hydraulics 109


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

uin++1/1 2, j - uin+1/ 2, j V in++1,qj - V in, +j q


+ a11 ( u n , u n ) + a12 ( v n , u n ) + g +
Dt Dx

( )
2
n +1 æ n ö 2 (6.14)
ç ui +1/ 2, j ÷ + vi +1/ 2, j
n
u i +1/ 2, j
è ø
cf * n
=0
hi +1/ 2, j

vin, +j 1+1/ 2 - vin, j +1/ 2 V in, +j +q1 - V in, +j q


+ a21 ( u , v n n
) + a22 ( v , v ) + g
n n
+
Dt Dy

( )
2
æ ö
2
(6.15)
n +1
v
i , j +1/ 2
n
u
i , j +1/ 2 + ç vin, j +1/ 2 ÷
è ø
cf * n
=0
hi , j +1/ 2

where the symbol * has the same meaning as in Equation (5.18) again and a11(un,un),
a12(vn,un), a21(un,vn) and a22(vn,vn) are generalizations of the discretization of the
convective momentum term.
The long double bar over the velocity in the friction term means that this velocity is
obtained by averaging over values at four surrounding grid points. The friction term
requires special treatment in case of flooding of dry terrain. At the wave front the
water velocity rapidly accelerates from zero. Overshoot of velocities can be
prevented by a predictor corrector approach.

The convective momentum terms are subject to the same principles as discussed for
the 1D approximations. For example, for positive flow velocities the momentum
conservative discretization of the term a12(un,vn) is given by

x
q1+1/ 2, j -1/ 2 æ ui +1/ 2, j - ui +1/ 2, j -1 ö
a12 (v , u ) ;
n n v
ç ÷ (6.16)
x
hi +1/ 2, j è Dy ø
whereas it is given by
x æ ui +1/ 2, j - ui +1/ 2, j -1 ö
a12 (v n , u n ) ; v i +1/ 2, j -1/ 2 ç ÷ (6.17)
è Dy ø
x
for the energy conservative discretization. In the first expression v q means the
specific discharge in y-direction, averaged over two surrounding points along the
x
local x-axis. In the last expression v has the same meaning in relation to the
velocity v.

The treatment of the convective momentum terms shown above is numerically very
robust and allows for the correct description of the effects of sudden expansions and
contractions and similar changes in the topography, such as steps in the bed level.
Moreover, it allows for the 2D simulation of supercritical flows and the propagation
of hydraulic jumps.

WL | Delft Hydraulics 110


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Figure 6.5 Layout of the instantaneous dam break experiment

Figure 6.6 Top view of results of the numerical scheme of Equations (6.13), (6.14) and (6.15) (lower half),
compared with video monitored measurements in a physical model (upper half)

As described by Stelling and Duinmeijer (2003), the correct modelling of these


phenomena has been demonstrated in a software validation study, where results of
the numerical scheme of Equations (6.13), (6.14) and (6.15) were compared with
video monitored measurements in a physical model (Figure 6.6). The set-up consists
of two reservoirs with different water levels, separated by a wall. The wall contains
a gate which can be lifted. The width of both reservoirs is 8.30 m, the length of the

WL | Delft Hydraulics 111


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

upper reservoir 2 m and the length of the lower reservoir 29 m. The gate has a width
of 0.4 m and has been placed at the middle of the wall.

The numerical experiment was made on a grid of 0.1*0.1 m, giving a total of


approximately 25000 grid cells. The time step was set to 0.005 s. The gate was lifted
at a speed of 0.16 m/s, to produce a flow spreading out into a 2-dimensional plain.
Initial data were set at a depth of 0.60 m for the upstream reservoir and a depth of
0.05 m for the downstream reservoir.

Figure 6.7 Comparison of measured and computed wave front position at various times (see case Figure 4)

Figure 4 shows the results of this simulation. The upper half of this figure presents a
video recorded view from above. The lower part of the figure presents the computed
results. It is clearly seen that the front propagation, the propagation of the hydraulic
jump and the side spreading of the wave are represented reasonably well. Figure 5
shows a comparison of the measured and computed position of the wave front at
various times.

Simulations were made with various Manning roughness coefficients and both for an
initially wet and a dry downstream reservoir. For the propagation of the flood on the
dry bed the Manning roughness turned out to be a sensitive parameter. If for dam
break models a reliable topography is available, the roughness parameter remains
the only parameter to be estimated. Currently, researchers are focusing on better
descriptions of roughness parameters by deriving depth dependent relationships on

WL | Delft Hydraulics 112


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

the basis of vegetation characteristics, e.g. Uittenbogaard (2003), Rodriguez (2004)


and Baptist (2005).

6.8 Integrated 1D/2D modelling


In flood modelling, there are numerous practical examples where flows are best
described by combinations of 1D and 2D schematizations. An obvious example is
the flooding of deltaic areas, often characterized by a flat topography with complex
networks of natural levees, polder dikes, drainage channels, elevated roads and
railways and a large variety of hydraulic structures. Flow over the terrain is best
described by the 2D equations, whereas channel flow and the role of hydraulic
structures are satisfactorily described in 1D. Flow over higher elevated line
elements, such as roads and embankments can be reasonably modeled in 2D by
raising the bottom of computational cells to embankment level. However, for a
higher accuracy of the numerical description adapted formulations have to be
applied, such as energy conservation upstream of overtopped embankments.

Another example is the flood propagation in a meandering river, with shortcuts via
the flood plain when over bank flow occurs. In large scale models, the flow between
the river banks is satisfactorily described by the de Saint Venant equations solved
with 1D grid steps several times the width of the channel. An equivalent accuracy of
description in 2D would require a large number of grid cells, with step sizes being a
fraction of the channel width. However, flow in the flood plain may be better
described in 2D and may allow for 2D grid steps often exceeding the width of the
river.

For this reason, hybrid 1D and 2D schematizations are often used. Basically there
are two approaches: one with interfaces defined between 1D and 2D along vertical
planes and the other approach with schematization interfaces in almost horizontal
planes.

Coupling along vertical planes, gives a full separation in the horizontal space of the
1D and 2D modeled domains. In the 1D domain the flow is modeled with the de
Saint Venant equations applied over the full water depth. The direction of flow in
the 1D domain is assumed to follow the channel x-axis and in the model it carries its
momentum in this direction, also above bank level. Without special provisions, there
is no momentum transfer accounting applied between the 1D and 2D domains.
Momentum and volume entering or leaving the 2D domain at these interfaces, are
generated by the compatibility condition applied. As a result, the coupling cannot be
expected to be momentum conservative. Depending on the numerical solution
applied, the linkage may either be on water level or on discharge compatibility.
Particular care has to be taken in applying this form of schematization if water
quality processes are to be included in the model.
In a model coupled along an almost horizontal plane, 2D grid cells are placed above
the 1D domain, as shown in Figure 6. In this schematization, the de Saint Venant
equations are applied only up to bank level. Above this level, the flow description in
the 2D cell takes over. For relatively small channel widths compared to the 2D cell
size, errors in neglecting the effect of momentum transfer at the interface are minor.
For wider channels it is recommended to modify each 2D cell depth used in the

WL | Delft Hydraulics 113


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

momentum equation by adding a layer defined by the local hydraulic radius for that
part of the 1D cross-section which underlies a 2D cell.
Further refinements are possible, including terms describing the momentum transfer
between the 1D and 2D domains.

Numerical solutions are obtained by discretizing separately the 1D and 2D domains.


Assuming that for both domains implicit numerical schemes are applied, the
interface compatibility conditions can be modeled either as an explicit or an implicit
link. Applying explicit links, first the solutions for the 1D and 2D domains are
generated sequentially. Subsequently, exchange flows are computed and added as
lateral flows at the next time step. Implicit links are based upon water level
compatibility. These equations are then added to the complete sets of equations
generated separately for the 1D and 2D domains. There are many approaches to
solving the complete set of equations. With the current state of the art, it is no longer
necessary to apply for the 1D domain different solvers for so-called simply or
multiply connected channel networks. Similarly, in 2D there is no real need anymore
for alternating direction algorithms, as the efficiency of the conjugate gradient
solvers has increased significantly over the past years.

Figure 6.8 Coupling of 1D and 2D domains in SOBEK

As an example, Delft Hydraulics has developed its combined 1D2D package


SOBEK for the modelling of integrated fresh water systems (www.sobek.nl;
www.wldelft.nl). The 1D part of hybrid models is based upon the numerical scheme
of Equation (5.20) and (5.31). The 2D part is described by the single step 2D scheme
given by Equations (6.13), (6.14) and (6.15). For efficiency reasons, the continuity
equations for the 1D and 2D domains are combined into one single equation at
points where 1D grid sections underlie a 2D cell. As a first step in reducing the total
number of equations, SOBEK eliminates all equations at velocity grid points. The
second step in the solution algorithm is the elimination of a large number of
unknowns by applying a minimum connection search between unknown water
levels. As a rule, this leads to an efficient elimination of nearly all unknowns of the
1D domain and a substantial number of unknowns in the 2D domain.

WL | Delft Hydraulics 114


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

This direct solver carries its elimination on, until nearly every second equation in the
2D domain has been eliminated. Beyond this point, it is more economical to apply
the conjugate gradient solver to solve the remaining set of equations.

Apart from its efficiency, an additional advantage of eliminating nearly every


second 2D equation is the improved conditioning of the resulting matrix. This
follows from the fact that elimination of an unknown water level at a 2D grid point
has the effect of increasing the spatial distance between the remaining adjacent
points, where water levels are still unknown. This, in turn, reduces Courant numbers
and as a consequence, it leads to changed coefficients at the main diagonal of the
matrix which is now more dominant in relation to the other diagonals, e.g. Verwey
(1994).

Figure 6.9 Flood modelling of the Vallei and Eem area, The Netherlands

An example of a combined 1D2D model is shown in Figure 7. It represents the


schematization of a model of the Eem Valley area in The Netherlands applied in a
study of the potential effects of a River Rhine dike breach.
This model has been used to provide information on warning lead times and flood
depths for evacuation planning. The Rhine branch upstream of the breach has been
modeled in 1D. At its upstream end a design hydrograph was specified, whereas the
downstream boundary condition of this relatively short branch is given by a rating
curve. Such a short downstream reach is permissible as the rating curve
automatically corrects for most of the effects of flow deviated through the breach at
this boundary. The breach itself has been described in 1D as a structure with a
velocity dependent breach growth.

WL | Delft Hydraulics 115


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

North of the dike the 1D link discharges into the 2D domain, given by a 100 * 100
m grid with bottom levels derived from a digital elevation model and resistance
coefficients derived from land use maps. Elevated roads and railways are presented
as flow barriers by raising the underlying cell bottom levels up to the levels of these
embankments. The resulting flood depths presented in Figure 7 clearly show the
effect of the 1D channel in the schematization. Due to their greater depth, flood
waves propagate faster in these channels than over land. Further downstream, this
leads to first signs of the progressing flood wave already one or two days before the
main flood arrives.

6.9 Exercise
At a discharge measuring station a cross-section is measured as shown below.

a) Compute the conveyance of the channel for levels h=2.00 m and h=3.00 m,
respectively. The Manning numbers vary along the cross-section and their
values have been indicated in Figure 6.8..
b) Compute discharges for both levels given an approximate bed slope I0=10-3.
c) Compute approximate flood wave celerities for the given water levels. Explain
the difference in the celerities.
d) During the passage of a flood wave, the water level rises from 2.00 m to 3.00
m above the reference level over a period of 10 minutes. Compute the
approximate deviation between the water level slope and the bed slope. (Note:
transform the time to a length via the wave celerity).
e) Compute the approximate difference in percentage between the expected
measured discharges and the discharges based on a rating curve.

Figure 6.10 Cross-section for the computation of flood wave propagation celerities

WL | Delft Hydraulics 116


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

7 Water Hammer

7.1 Introduction
Water hammer is the phenomenon of high-amplitude pressure waves travelling in
pipes, caused by the rapid changes in the velocity of the transported fluid. These
variations usually result from the operation of flow- and pressure control devices,
such as valves, pumps and turbines.

d slot

Figure 7.1 Sketch of pipe with artificial slot at the top to represent storage effects of pipe expansion
and water compressibility

In principle, the equations used for the computation of these pressure waves are the
same as those derived for open channel flow. As an analogy one may consider a pipe
which has an artificial slot at the top, in which a free surface water level may rise
after complete filling of the pipe (Figure 7.1).

The wetted area of the slot represents the storage of water resulting from changes in
the water pressure. The storage capacity consists of three principal contributions:

1. the elasticity of the pipe wall which leads to a pipe cross-section expansion at
increasing water pressure. It is evident that this type of storage depends on the
cross-section shape and on the wall material properties and its composition.
2. compressibility of the water, which usually is neglected in free surface flow
computations. In the case of water hammer, however, the compressibility plays a
dominant role. As the free surface flow equations are based upon a volume
balance, the compression of the water represents a virtual water storage;
3. compressibility of air bubbles or vacuum bubbles contained in the water;

Water hammer is important in engineering for the following reasons:

1. high pressures may build up in a pipe line to the extent that it may lead to
bursting of the pipe;

WL | Delft Hydraulics 117


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

2. vacuum bubbles may be generated when the pressure drops to values below
approximately 30% of the atmospheric pressure. The resulting cavitation is
highly aggressive to the pipe material and may lead to considerable damage;
3. in other cases, air may be entrained in the fluid at moments and locations where
the pressures are low.

For these reasons, any design related to transport of fluid in pipes should be checked
against the risk of water hammer damage. There are various reasons why water
hammer may occur. For details, reference is made to text books, e.g. Chaudhry
(1987). Typical reasons are:

· instantaneous or very fast valve closure. Many of us know this phenomenon


when closing a tab at home. A wrong design of the valve may cause a sudden
deceleration of the water flow and a high pressure builds up due to the
transformation of momentum of the water into an impulse;
· sudden energy demand changes in a hydropower station or turbine failure;
· pump failure, for example, by a power cut. At the upstream end of the pump the
pressure will increase, while at the downstream end a negative wave is generated
which may cause cavitation or implosion of the pipe;
· burst of a pipe line, which in turn generates pressure surges along the pipe.

In many cases it is the way of operating the system that causes the water hammer
problems. In other cases there may be accidental causes. In the design of the
hydraulic system such operation problems may be foreseen and water hammer may
be prevented by designing anti-water hammer arrangements. Examples are:

· strict control of valve manipulations. As will be shown by the computational


procedure, slow closure of valves reduce the over pressures. Slow has to be seen
in relation to the length of the pipe and the celerity with which pressure surges
are travelling along the pipe;
· design and construction of a surge chamber as closely as possible upstream of
the turbine or pump. This surge chamber serves as an escape for the pressure
wave and reduces effectively the dangerous operation time of closure;
· closed and vented air vessels, which have a function similar to surge chambers.
However, in this case the surface is formed by a pressurized air chamber. These
devices are used, for example, at the downstream end of sewer pumps in order to
prevent excessive under pressures;
· fly wheels, which prevent the sudden change of the pump rotational speed;
· by-passes with a check-valve placed in parallel to a pump. It opens during pump
rundown and supplies water to the pipe downstream of the pump;
· pressure release valves, which open when the water pressure exceeds a
maximum value admitted.

For a more extensive list of options reference is made again to the literature in this
field. As water hammer computations may be complex, use is often made of
standard software packages, such as the WANDA system of Delft Hydraulics
(www.wldelft.nl). However, simple problems may also be investigated by using
Excel. The next paragraphs provide the basis for setting up such computations.

WL | Delft Hydraulics 118


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

7.2 Water Hammer Equations


For a discussion on the water hammer equations reference is made to the open channel
flow equations (Chapter 5), given here as

¶H ¶Q
bs + = 0 (7.1)
¶t ¶x

¶Q ¶ æ Q 2 ö ¶H f
+ ç ÷ + gA + Q|Q| = 0 (7.2)
¶t ¶x è A ø ¶x 2DA

where

Q = pipe discharge (m3/s);


H = piezometric head above any horizontal reference level (m);
bs = storage width due to fluid compressibility and pipe expansion (m);
A = cross-sectional area of the pipe (m2);
f = Darcy-Weisbach friction factor;
D = pipe diameter (m).

In the momentum equation the friction term has been replaced by a term based on the
Darcy-Weisbach concept. In this form, the coefficient expresses the energy head loss
DH as the fraction of the velocity head which is lost over a pipe length equivalent to
its diameter D. For a length L of the pipe the head loss is then defined as

L u2
DH = f (7.3)
D 2g

In water hammer, the role of the convective momentum is small and usually
neglected. Introducing, furthermore, a characteristic celerity a as

A
a = g (7.4)
bs

the water hammer equations are best known in the form

gA ¶H ¶Q
+ = 0 (7.5)
a ¶t ¶x
2

¶Q ¶H f
+ gA + Q|Q| = 0 (7.6)
¶t ¶x 2DA

The physical behaviour of water hammer is easily understood by transforming these


equations into the form

WL | Delft Hydraulics 119


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

¶ æ gA ö ¶Q
ç H÷ + a = 0 (7.7)
¶t è a ø ¶x

¶Q ¶ æ gA ö
+a ç H÷ = E (7.8)
¶t ¶x è a ø

where E is to be seen as a sink term representing the effect of friction losses.

f
E = - Q|Q| (7.9)
2DA

Successive addition and subtraction of Equations (7.8) and (7.7) gives

¶ æ gA ö ¶ æ gA ö
çQ + H÷ + a çQ + H÷ = E (7.10)
¶t è a ø ¶x è a ø

¶ æ gA ö ¶ æ gA ö
çQ - H÷ -a çQ - H÷ = E (7.11)
¶t è a ø ¶x è a ø

Referring to the concept of total derivatives, as introduced in Chapter 3, Equations


(7.10, 7.11) express the integration of the equation

d æ gA ö
çQ + H÷ = E (7.12)
dt è a ø

along the characteristic line, or simply characteristic, defined by

dx
= a (7.13)
dt

and the integration of

d æ gA ö
çQ - H÷ = E (7.14)
dt è a ø

along the characteristic line

dx
= -a (7.15)
dt

Neglecting the influences of friction, Equation (7.12) represents the condition

gA
Q+ H = constant (7.16)
a

WL | Delft Hydraulics 120


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

along the positive characteristic given by Equation (7.13), while Equation (7.14) leads
to

gA
Q- H = constant (7.17)
a

along the negative characteristic, given by Equation (7.15).

Before discussing the role of these equations in the algorithm leading to the solution
of water hammer problems, let us consider again the meaning of the characteristic
celerity a in relation to water storage. In analogy with Hooke's law expressing the
elastic deformation of a string under the influence of the axial stress

s = eE (7.18)

where
s = axial stress (N/m2);
e = strain per unit length of the string;
E = Young's modulus of elasticity.

The compression of water under the influence of changing pressures is seen as an


elastic process described by

dV
dp = - K (7.19)
V

where

dp = change in fluid pressure (N/m2);


K = bulk modulus of water elasticity (N/m2);
V = water volume;
dV = change of this water volume under the influence of pressure change dp.

For a control volume of length dx the compression of fluid in the pipe can be made
equivalent to a virtual storage of an incompressible fluid in a virtual pipe slot by the
relation

V
dV = - dp = - b s dH dx (7.20)
K

where the change in pressure level dH is related to the change in water pressure dp, by

dp = r g dH (7.21)

assuming that the contribution of the change in fluid density in this relation can be
neglected.

WL | Delft Hydraulics 121


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Substitution of (7.21) into (7.20) and replacing dV by Adx, gives

rgA
bs = (7.22)
K

For the derivation of the effect of pipe deformation it will be assumed that the pipe
wall is elastic and that the pipe has expansion joints all along its axis. These joints
will prevent the development of axial stresses under the influence of changing water
pressure in the pipe. The general stress-strain relation for elastic pipe material under
the influence of stresses in axial and tangential direction reads as

s 1 = ms 2 = e E (7.23)

where
s1 = hoop stress (tangential direction) (N/m2);
s2 = axial stress (N/m2);
m = Poisson's ratio.

For the case where the axial stress cannot develop the equation reads, for a change in
tangential stress

ds 1 = e E (7.24)

The expression of e in terms of the change in the circular pipe circumference with
radius r

d(2p r) ds 1
e = = (7.25)
2p r E

now leads to the relation

dr ds 1
= (7.26)
r E

or

dA 2ds 1
= (7.27)
A E

Referring to Figure (7.2) the change in pipe material stress is related to the change in
pressure dp as

2dF = Ddp (7.28)

or

WL | Delft Hydraulics 122


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

2eds 1 = Ddp (7.29)

P
e

t t

Figure 7.2 Forces on semi-cylinder under the influence of water hammer induced pressures

Substitution of Equations (7.29), (7.21) into Equation (7.27) gives

D
dA = r gA dH (7.30)
eE

Relating this compression volume to the water storage in the fictive pipe slot gives

D
dV = dA dx = rgA dH dx = b s dH dx
e

or
D
bs = r g A (7.31)
eE

The combined effects of water compressibility and pipe expansion leads to the virtual
pipe slot width

r g A æ DK ö
bs = ç1 + ÷ (7.32)
K è eE ø

Substitution of this relation into Equation (7.41) gives the expression for the pressure
wave celerity

K
r
a = (7.33)
DK
1+
eE

The effect of axial stresses and composite pipe walls requires a further generalisation
of this equation to the form

WL | Delft Hydraulics 123


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

K
r
a = (7.34)
K
1 +y
E

with

D
y = (7.35)
e

for the case of homogeneous elastic pipe material and the presence of expansion joints
all along the pipe.

Referring to Chaudhry (1987), the following relations can be derived for Y:

D
y = (1 - m 2) (7.36)
e

for the case of thin-walled elastic conduits anchored against axial movement
throughout their length;

D
y = (1 - 0.5m ) (7.37)
e

for the case of thin-walled elastic conduits anchored against axial movement at the
upper end;

y =1 ; E=G (7.38)

for the case of an unlined tunnel with a modulus of rock elasticity G, and

DE
y = (7.39)
GD + Ee

for a steel-lined tunnel through rock.

For values of the various material properties reference is made to literature and to
Table 7.1 and Table 7.2 for some very common ones.

WL | Delft Hydraulics 124


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Table 7.1 Pipe material properties

Material Modulus of elasticity E Poisson’s ratio


(Gpa)

Mild steel 200 - 212 0.27


Polyethylene 0.8 0.46
Concrete 14 –30 0.10 –0.15
Cast iron 80 –170 0.25

Table 7.2 Fluid properties

Fluid Temperature Density Bulk modulus of elasticity K


(oC) (kg/m3) (GPa)

Fresh water 20 999 2.19


Sea water 15 1025 2.27
Oil 15 900 1.5

7.3 The Method of Characteristics for Water Hammer


The method of characteristics is by far the most commonly used equation solver in
water hammer analysis. In the discussion of Chapter 3 it was concluded that the
method of characteristics is a rather unpractical approach to solving free surface
flow problems. Most of the reasons given there, however, do not hold for water
hammer analysis. As shown by Equations (7.13, 7.15 and 7.33), the characteristic
directions for the water hammer equations usually are straight lines, which allow for
the construction of an equidistant network, as shown by the following example.

Consider the hypothetical case of a pipe with frictionless flow supplied at the
upstream end from a large reservoir and controlled at the downstream end by a
valve, as shown in Figure (7.3). Data for this problem are given as follows:

Hres = 50 m
Q = 1 m3/s
A = 1 m2
L = 1000 m
a = 1000 m/s
g = 10 m/s2

The valve is closed over a period of 5 seconds, controlled to provide a linearly


decreasing discharge over this time.

WL | Delft Hydraulics 125


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

The characteristic grid for this problem is shown in Figure 7.4. Initially all
discharges and pressure levels are equal to 1 m3/s and 50 m respectively.

Referring, again, to Equations (7.16 and 7.17), the solution at point A1 is found by
solving the boundary condition

H A1 = 50 m

and by Equation (7.17) applied along the negative characteristic from point B0 to point
A1

Q A1 - gA/a H A1 = Q B0 - gA/a H B0 = 1 - 0.01*50 = 0.50 m3 /s

giving

H A1 = 50 m; Q A1 = 1.0 m 3 /s

Note that this solution is still equivalent to the initial steady state condition along the
pipeline between A and B. This is explained by the fact that the effect of the valve
closure starting at point B0 cannot influence conditions at the upstream boundary
before the negative characteristic through point B0 has arrived at point A1. The
triangle formed by the points A0, B0 and A1, therefore, is a steady state region.

The solution at point B1 is found in a similar way by combining the downstream


boundary condition at t=1 s, given as

Q B1 = 0.8 m3/s

with the condition along the positive characteristic passing from A0 to B1, given as

Q B1 + gA/a H B1 = Q A0 + gA/a H A0 = 1 + 0.01*50 = 1.50 m3 /s

resulting in

H B1 = 70 m; Q B1 = 0.8 m3 /s

Similarly, solutions at subsequent points can be computed, with results shown in


Table 7.3. The maximum pressure found at point B is 90 m water column.

The gradual valve closure gives a considerable operation improvement over the case
of a sudden valve closure. Closing the valve instantaneously at time t=0 gives a
pressure H=150 m at point B1, as can be verified easily by applying Equation (7.16)
with QB1=0 m3/s. Further improvements are found by slowing the closure operation
further down. This example, therefore, demonstrates clearly the origin of water
hammer problems and the essential approach to water hammer prevention.

WL | Delft Hydraulics 126


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

reservoir

supply pipe

Figure 7.3 Situation sketch of valve closure problem


t (s)

t (s)
10 10

9 9

8 8

7 7

6 6

5 5

4 4 0.2

3 3 0.4

2 2 0.6

1 1 0.8

0 0 1.0
0 100 0 1.0

x (m) Q (m 3 /s)

A B

Figure 7.4 Characteristic lines for the water hammer problem

Table 7.3 Pressures and discharges computed for the problem of valve closure

Point 0 1 2 3 4 5 6 7 8 9 10 variable

A 1.0 1.0 0.6 0.2 0.2 0.2 -0.2 -0.2 0.2 0.2 -0.2 Q (m3/s)
50 50 50 50 50 50 50 50 50 50 50 H (m)
B 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.0 0.0 0.0 0.0 Q (m3/s)
50 70 90 70 50 70 70 30 30 70 70 H (m)

WL | Delft Hydraulics 127


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

7.4 Exercise
Water flows from a reservoir through a horizontal pipe with a length L=10 km. The
reservoir water level HR is 50 m +MSL. The pipe has a diameter D of 1.00 meter. At
the end of the pipe the outflow is controlled by a valve. The steady state water velocity
u=2 m/s. The Darcy-Weisbach friction coefficient f is 0.01. Use g=10 m/s2 in your
computations.

Questions

1. Compute the pressure distribution along the pipe line for the steady state flow,
assuming that the energy loss at the valve can be neglected.
2. For the computation of Questions 3-7 we will assume that the complete energy
loss along the pipe is now concentrated at the valve. Recompute the pressure
distribution.
3. Compute the characteristic pressure wave celerity for the data given below.
4. Compute, with the rigid water column theory, the pressure increase at the pipe
end if the valve is closed over a period of 1 second. Is this pressure increase
realistic?
5. Compute the pressure increase with the water hammer theory for fast closure.
6. Compute the pressure as a function of time at the location of the valve by
using the method of characteristics. Use a valve closure function which leads
to a linear decrease in out flowing discharge over a period of 1 minute. Neglect
friction.
7. While keeping the other parameters constant, what will be the effect on the
maximum pressure increase if:
a) the closure time is increased;
b) the steel pipe is replaced by a PVC pipe;
c) the pipe wall thickness is increased;
d) the pipe diameter is increased;
e) the pipe length is decreased;
f) air is entrained into the flow.
8. Repeat the computations for the case where pipe wall friction is included and
the energy loss at the valve is neglected. Produce a graph of the results.

Additional data:

e = 0.01 m;
K = 2 GPa;
E = 200 GPa.

The pipe has frequent expansion joints.

WL | Delft Hydraulics 128


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

8 References
Abbott, M.B. (1979). Computational Hydraulics – Elements of the Theory of Free Surface
Flows, Pitman, London/San Francisco/Melbourne.

Arakawa, A. (1966). Computational design for long-term numerical integration of the


equations of fluid motion: two-dimensional incompressible flow, Part I. Journal of
Computational Physics. 1(1), 119-143.

Baptist. M.J. (2005). Modelling floodplain biogeomorphology. Ph.D. thesis, ISBN


90-407-2582-9, 193 pp., Delft University of Technology, Faculty of Civil
Engineering and Geosciences, Section Hydraulic Engineering.
Chaudhry, M.H. (1987), Applied Hydraulic Transients, Van Nostrand Reinhold Company
Inc., New York.

Chaudhry, M.H. (1993), Open-Channel Flow, Prentice Hall, Anglewood Cliffs, USA.

Chow, V.T. (1959). Open Channel Hydraulics, McGraw-Hill, New York.

Chanson, H. (1999). The Hydraulics of Open Channel Flow, Arnold Publishers/Wiley,


Paris/New York.

Cunge, J.A., (1969). On the subject of a flood propagation computation method


(Muskingum method). Journal Hydrology Research 7, No.2.

Cunge, J.A., Holly, F.M. and Verwey, A. (1986). Practical Aspects of Computational River
Hydraulics (420 pages), Pitman Publishing Ltd., London, Great Britain, 1980. Reprinted at
Iowa Institute of Hydraulic Research, USA, 1986. (Also translated into Russian).

De Saint Venant, B. (1871). Théorie du Mouvement Non-Permanent des Eaux avec


Application aux Crues des Rivières et à l’Introduction des Marées dans leur Lit, Comptes
Rendus de l’Académie des Sciences, 73, pp. 148-154, 237-240.

Henderson, F.M. (1966). Open Channel Flow, McMillan Company, New York.

Hesselink, A.W., Stelling, G.S., Kwadijk J.C.J. and Middelkoop, H. (2003). Inundation of a
Dutch river polder, sensitivity analysis of a physically based inundation model using historic
data. Water Resour. Res., 39(9), 1234.

Hirsch, C. (1990). Numerical Computation of Internal and External Flows, Wiley, New York.

Newton, I. (1687). Mathematical Principles of Natural Philosophy (trans. A. Motte, ed. F.


Cajuri), Berkeley University Press, California (1947).

Rodriguez Uthurburu, R, (2004). Evaluation of physically based and evolutionary data mining
approaches for modelling resistance due to vegetation in SOBEK 1D-2D, M.Sc. thesis HH
485, UNESCO-IHE, Delft, The Netherlands.

Serban, P., Crookshank, N.L. and Willis, D.H. (2005). Intercomparison of Forecast Models
for Streamflow Routing in Large Rivers, WMO, Geneva.

WL | Delft Hydraulics 129


Computational Hydraulics © 2005 Adri Verwey UNESCO - IHE
Version 2006-1

Stelling, G.S., Kernkamp, H.W.J. and Laguzzi, M.M. (1998). Delft Flooding System: a
powerful tool for inundation assessment based upon a positive flow simulation,
Hydroinformatics ‘98, Babovi & Larsen (Eds.), Balkema, Rotterdam, 449-456.

Stelling, G.S. and Duinmeijer, S.P.A. (2003). A staggered conservative scheme for every
Froude number in rapidly varied shallow water flows. Int. J. for Numer. Meth. Fluids, 43,
1329-1354.

Stoker, J.J. (1957). Water Waves, Pure and Applied Mathematics, Vol. IV, Interscience
Publishers, New York.

Toro, E.F. (1999). Riemann Solvers and Numerical methods for Fluid Dynamics, Springer,
Berlin.

Uittenbogaard, R. (2003). Modelling turbulence in vegetated aquatic flows,


presented at the International workshop on RIParian FORest vegetated channels:
hydraulic, morphological and ecological aspects, 20-22 February 2003, Trento,
Italy.
Verwey, A. (1994). Linkage of Physical and Numerical Aspects of Models Applied in
Environmental Studies, keynote lecture in: Proceedings of the Conf. on Hydraulics in Civil
Engineering, Brisbane, Australia.

Verwey, A. (2001). Latest Developments in Floodplain Modelling - 1D/2D Integration,


keynote lecture in : Proceedings of the 6th Conf. on Hydraulics in Civil Engineering,
Hobart, Australia.

Werner, M.G.F., van Dijk, M. and Schellekens, J., (2004). DELFT-FEWS: An


open shell flood forecasting system, in Proceedings of the 6th International
Conference on Hydroinformatics, Liong, Phoon and Babovi (Eds.), World
Scientific Publishing Company, Singapore, 1205-1212.

WL | Delft Hydraulics 130

Anda mungkin juga menyukai