ON THE RESOLUTION OF THE NAVIER-STOKES EQUATIONS BY THE
FINITE ELEMENT METHOD USING A SUPG STABILIZATION
TECHNIQUE. Application to some wastewater treatment problems
*P. Vellando, **J. Puertas, *I. Colominas
ETS. de Ingenieros de Caminos, Canales y Puertos.
*Dpto. de Métodos Matemáticos y de Representación.
**Dpto. de Tecnología de la Construcción.
Universidad de La Coruña. Campus de Elviña. 15071 La Coruña, Spain.
Tel: (34) 981 167000, fax: (34) 981 167170. e-mail: vellando@iccp.udc.es
ABSTRACT
In this paper an analysis of the incompressible flow has been carried out, from the very
definition of the governing equations, up to the resolution of some practical problems,
passing through the comprehensive study of the stabilized finite element techniques
used in their resolution. As a consequence of this analysis, a code based upon a realistic
interpretation of the forces has been written, which allows for the modelling of the open
channel flow, with optimum results in the resolution of some benchmark and real flow
problems related with the wastewater industry.
Key words: Viscous incompressible flow, FEM, Navier Stokes, SUPG.
INTRODUCTION
The Finite Element Method was first developed in the fifties by Turner and Clough so
as to solve some structural problems of the aeronautical industry [Turner 56]. The
application of the Finite Element Method to the flow problems requires some
modifications with respect to the formulation used for the structural stress analysis
problems that were its first application. Some of these modifications have been
borrowed from the finite difference or finite volume approaches, and many others have
been specifically developed for finite elements. In the early seventies we find many
works regarding not only the mere existence and consistency of these flow problems
[Ladyzhenskaya 69], [Babuska 71], [Brezzi 74], but also many works that give a finite
element solution to the Navier-Stokes equations [Baker 71], [Oden 72], [Fortin 72],
[Taylor 73], [Shen 76], [Zienkiewicz 76]. Since then, the Finite Element Method is a
powerful tool for the resolution of the Navier Stokes equations, which will be used in
this work to solve the incompressible flow.
The material we are going to deal with, when solving the flow, is of a fluid nature, and
therefore it has not a fixed shape, which is instead a function of time. In addition to
Newton’s second law, that rules any dynamic problem, an equation that ensures for the
conservation of mass should be verified. Moreover, the Navier-Stokes equations are a
set of differential equations with respect to both space and time in which both the
pressure and the velocity are the unknowns. As a consequence, the finite element
formulation used for the conventional structural analysis cannot be applied
straightforwardly.
When applying the finite element analysis to the problems of the rigid body, the
weighted residual method can be exclusively applied to the Newton second law, which
1
for statics clearly turns out to be the equilibrium equation. On the contrary, when
dealing with fluids, the shape is not any more conserved, and apart from stating the
equilibrium of momentum, we have to ensure for the continuity of mass. Consequently,
we have two equations to be verified at the same time, and the finite element
formulation should also account for the verification of both. The only set of unknowns
in the conventional structural analysis is that of the displacements, as a consequence, the
system obtained thanks to the application of the Finite Element Method, gives the
displacements in the structure depending on the stiffness matrix (that features the
structure), and the load vector. In the flow problems, we are headed towards the socalled mixed Finite Element Methods, in which both the velocity and pressure set of
unknowns have to be treated simultaneously.
Depending on how these two sets of equations and unknowns are tackled, several
different approaches have been developed. Some authors [Kim 88], [Choi 94], agree to
classify the different approaches to solve the viscous incompressible flow by the Finite
Element Method, into three different categories, these are the mixed (or velocitypressure integrated), the penalty and the segregated methods. The mixed method is the
most intuitive of these approaches, and is based upon carrying out a similar analysis for
the continuity equation to that used for the momentum equation, keeping both velocity
and pressure as the unknowns up to the end of the problem, [Baker 71] [Oden 72],
[Zienkiewicz 76]. This apparently straightforward way of dealing with our equations is
not as simple as it appears to be, and it may be the reason of the obtaining of a
meaningless solution when used in connection with a faulty basic element [Babuska
71], [Taylor 73], [Brezzi 74]. Besides a big expense in the storing memory, the socalled mixed formulation, leads to some consistency problems in the obtaining of the
solution when a wrong choice in the basic functions has been made. As a consequence,
the penalty and segregated methods have been developed to overcome these difficulties.
In this work, an algorithm of each type will be developed, programmed, verified,
employed and discussed.
The 2D Navier-Stokes equations assume a flow that takes place on a two-dimensional
plane, and it is therefore laminar in that sense. A fully 3D resolution of the flow, apart
from leading us to high rate memory requirements, would involve some problems in the
modelling of the free water surface. The Shallow Water formulation has been
considered as a way of including the third dimension in the calculations, being able to
give a meaningful solution for flows in which the depth is small compared to the
horizontal dimension. The integration in depth of the 3D Navier-Stokes formulation
causes the dependence of the continuity equation with respect to depth, and
consequently the appearance of some quasi-non-linear terms that depend on both the
velocity and the depth. These equations are solved thanks to a newly developed iterative
algorithm, which will be solved on a mixed formulation basis to be described later in
the text.
The use of a Galerkin formulation, that takes weighting functions equal to trial
functions when solving the Navier-Stokes equations, may lead to some problems of
instability in the flow solution by the Finite Element Method. To avoid this difficulty,
some so-called stabilization procedures have been released since the MAFELAP
conference in 1975 [Zienkiewicz 76]. The stiffness matrix resulting from structural
problems solved by the Finite Element Method is symmetric, instead the ‘stiffness’
matrix obtained for fluids is non-symmetric and the use of symmetric weighting
functions may lead to some instability problems. The faster the flow turns, the more
non-symmetric the coefficient matrix becomes. In practice this is featured by the
appearance of some spurious node-to-node oscillations also known as ‘wiggles’. One
2
way of avoiding these oscillations is to carry out a refinement in the mesh, such that
convection no longer dominates on an element level, but this refinement turns to be a
memory resources sink. This point will be avoided in this work by the use of a
stabilization technique of the SUPG type, for all the algorithms considered in it.
The SUPG (Streamline/Upwinding Petrov-Galerkin) technique, first developed by
Brookes [Brookes 82], succeeds in eliminating the spurious velocity field, without
carrying out a severe refinement in the mesh, by considering weighting functions that
differ from trial functions in an upwinding term. This method was first released for the
transport equation, and its generalisation to the Navier-Stokes equations brings an
additional problem; that is the appearance of an excessive diffusion normal to the flow.
The SUPG method eliminates this spurious crosswind diffusion by considering an
artificial diffusion that acts only in the direction of the flow.
All the particulars regarded in this introduction and some others, will be further
considered in the following chapters.
GOVERNING EQUATIONS
As in any other dynamic problem, the equation we are going to refer to, is the Newton
second law, which gives the variation in the momentum as the summation of the acting
forces on the volume of integration. To this condition we should add another one, due to
the fact that we are dealing with a shape-changing matter in which we have to ensure
the conservation of mass. Both equations make up the Navier-Stokes equations. These
relations are named after their discoverer, the French civil engineer Claude-Louis
Navier (1785-1836), who in 1821 formulated the equations that rule the incompressible
flow. The Navier-Stokes equations also bear the name of the Irish mathematician
George Gabriel Stokes (1819-1903), who not knowing the previous discoveries made
by Navier, Poisson and Saint-Venant, re-obtained the Navier-Stokes equations for
slightly different assumptions, and published these works in 1845. The Irish
mathematician gives also his name to the simplified version of the Navier-Stokes
equations, in which the convective terms are dropped.
Using the indicial notation in which the indices after commas will stand for derivatives
with respect to the variables specified in the index, we can express the Navier-Stokes
equations as:
1
ui ,t + u j ui , j = − p,i + νui , jj + f i
ρ
in
ui ,i = 0
together with the boundary and initial conditions:
σ ij n j
ui ]Γ1 = bi
ui (x j ,0 ) = ui 0 (x j )
]
Γ2
Ω
(1)
= ti
(2)
where u i is the velocity, p is the pressure, f i is the body force per unit mass, ρ is
density, ν is the cinematic viscosity, Γ1 and Γ2 are two non overlapping subsets of the
piecewise smooth domain boundary Γ , b i is the velocity vector prescribed in Γ1 , t i is
the traction vector prescribed on Γ2 , σ ij is the stress along the boundary Γ2 , and n j is
the outward unit vector normal to Γ2 .
The 2D Navier-Stokes equations will be used in this thesis to solve many benchmark
problems of the related literature with very good results, as will became clear later in the
3
text. The 2D or laminar (in the sense of planar) Navier-Stokes equations do not take into
account the third dimension in space, and provide with the velocities and pressures of a
theoretical planar flow. Nevertheless, for many real flow problems, the third dimension
in space is very important and the 3D Navier-Stokes equations should be considered.
The three-dimensional Navier-Stokes equations result in a very large-dimensioned
system of equations that involves very high computational costs. Moreover the 3D
schemes present a great difficulty in the treatment of the free surface. For flows in
which the horizontal dimension is small compared to depth, the Shallow Water
formulation can be employed as a simplification of the 3D Navier-Stokes equations
[Weiyan 92].
The Shallow Water equations are a simplification of the Navier-Stokes equations, which
can be used when the main direction of the flow is the horizontal one and the
distribution of the horizontal velocity along the vertical direction can be assumed as
uniform. These equations assume that the vertical acceleration of the fluid is negligible
and that a hydrostatic distribution of the pressure can be adopted. The Shallow Water
equations are obtained by integrating the 3D Navier-Stokes equations in depth, and give
a meaningful solution for flows in which the horizontal dimension is small compared
with the depth. When a 2D Navier-Stokes equation is used, no attention is paid to the
third dimension in space, and the results are based upon a 2D approach to the flow
problem. Therefore, the continuity equation is only held on a 2D basis. So as to get
some information about the variations in depth along the flow, either a 3D NavierStokes equation or the Shallow Water equations (if the flow can be regarded as
shallow), should be used. The Shallow Water equations are solved in this work for that
purpose. Integrating in depth the 3D Navier-Stokes equations, it is obtained:
h,t + hui ,i + ui h,i = 0
(3)
ui ,t + u j ui , j = − gh,i + νui , jj + g (S 0i − S fi )
where h is the depth, g is the gravity force and S 0i y S fi are the geometric and friction
slopes. Let us make some comments on the evaluation of the friction slope.
All the flows found in civil engineering practice can be featured by the Reynolds
number (UL/ν, where U and L are the characteristic velocity and length of the flow and
ν is the kinematic viscosity that depends on the fluid nature). For small Reynolds
numbers, the flow can be regarded as laminar, and the streamlines are parallel to each
other. As the Reynolds number is increased, a chaotic, random and intrinsically
unsteady type of motion appears. If these turbulent effects are to be solved by using the
Navier-Stokes equations, a very refined mesh would be required to capture the eddies
taking place on a wide range of length scales, and a special attention should be devoted
to the unsteady resolution of the turbulent phenomena, that take place at a very high
frequency [Versteeg 95]. The mesh refinement and the time step required for this
purpose are not yet computationally affordable and a turbulence model should be
implemented in order to evaluate these turbulent eddies. Most of these turbulence
models are based upon decomposing the involved variables into a mean value (within a
time increment) and a fluctuating term that depends on time. As a consequence of this
approach, a term that evaluates the turbulent losses as a function of a so-called eddy
viscosity ν t , is obtained. To evaluate this eddy or turbulent viscosity, a specific
turbulence model such as the k-ε model should be introduced. Making use of these
turbulence models, the turbulent viscosity is calculated for each time step and position,
allowing for the capturing of these eddies [Rodi 93]. Some other flow models evaluate
this eddy viscosity as a constant within the flow domain, such as the RMA2 flow model
4
developed by the Brigham University, which is one of the most commonly used
programs to evaluate the flow in channels.
Another approach to the turbulent problems would be to evaluate the friction slope on a
Manning basis. The integration in depth of the 3D Navier-Stokes equations allows for
the empirical evaluation of the energy losses taking place in flows that can be regarded
as shallow. The Manning formula evaluates empirically the overall energy losses taking
place in the fluid flow, including those related with the turbulent effects. This
formulation does not capture the turbulent eddies taking place within the fluid flow but
takes into account the overall turbulent energy losses, moreover keeps the full NavierStokes formulation, being ready to incorporate a k-ε model. By doing so, the frictional
slope can be taken as:
n 2 ui u 2j
(4)
S fi =
Rh4 / 3
where R h is the hydraulic radius and n is the Manning coefficient.
FINITE ELEMENT FORMULATION
Mixed laminar formulation
This formulation is directly obtained from the application of the weighted residuals
method on both the dynamic and continuity equations. In the Galerkin method, the
weighting functions are taken equal to the shape functions, in terms of which the
unknowns are interpolated. We are going to introduce an SUPG stabilization of the
algorithm by considering an additive term pi in the weighting functions. Applying the
weighted residuals method in this way on both the dynamic and continuity NavierStokes equations, we obtained:
1
h
h
h h
h
h
h
h
h h
∫ wi u i,t + u j u i, j − f i dΩ + ν ∫ wi, j u i, j dΩ − ∫ wi,i pdΩ − ∫ t i wi dΓ2 +
Ωh
(
)
ρ
Ωh
Ωh
1
h
+ ∑ ∫ pih uih,t + u hj ui,jh −νui,jj
+ p ,ih − f i h dΩ = 0 ;
ρ
e Ωe
∫q u
h
Ωh
h
i ,i
Γ2
dΩ = 0
(5)
where wi = wi + pi are the SUPG weighting functions for the dynamic equation and q
are the weighting functions for the continuity equation. The h superscript stands for the
discretization being carried out in our formulation, which will be made in terms of a
Q1P0 basic element:
uih = ∑ uij N i j
ph = ∑ p j χ j
1
4
j =1
j =1
(6)
This formulation should verify in addition to the governing equations, some consistency
conditions, the most restrictive of which is the LBB (Ladyzhenskaya-Babuska-Brezzi),
or divergence-stability condition [Babuska 71], [Brezzi 74]. A wrong election in this
basic element may fail to verify this condition and prevent the whole algorithm from
converging. The Q1P0 basic element, even not being strictly div-stability stable, has
shown not to produce the well known checkerboard pressure mode for the flow
5
problems to be considered in this work as will be shown later, providing with stable and
efficient solutions.
Once the elementary matrices are evaluated and assembled, the integral equation (5) can
be expressed in matrix notation as:
∂υ
Mυ
+ Cυ (u, v )υ + νAυ υ − Bp = f
∂t
(7)
BT υ = 0
where M υ is the mass matrix, Cυ (u, v ) is the convective matrix, A υ is the viscous
matrix, B is the pressure matrix, f is the external forces vector, p is the pressure vector,
u is the velocity vector in the x direction, v is the velocity vector in the y direction and
υ is the velocity vector.
For the derivatives with respect to time included in the first term of the first equation in
(7), a finite difference approach will be used in order to transform our partial differential
system into an algebraic one. This term can be evaluated on a backward differencing
scheme as:
υ n − υ n −1
∂υ
(8)
Mυ
≈ Mυ
t
∆
∂t
is the unknowns
where υ is the unknowns vector at the present iteration and υ
vector obtained in the former iteration. Therefore, the second term in expression (8) can
be brought to the right hand side of the equality.
The convective term Cυ (u, v )υ that appears in this formulation, is not the product of a
coefficient matrix times a vector of unknowns, but a non-linear velocity-dependent
function. This term should be eliminated in order to transform the resulting system into
a linear system of equations. The numerical scheme to be used for this transformation is
going to be a so-called successive approximation method, that can be mathematically
expressed as:
n −1 n
(9)
∫ wi u j ui, j dΩ ≈ ∫ wi u j ui, j dΩ
n −1
n
Ωh
Ωh
and has shown to provide with good results in the in the resolution of the Navier-Stokes
equations [Gartling 74]. This method converges linearly to the solution in opposition to
some others, like the Newton-Raphson method, which do it in a quadratic way in the
vicinity of the solution. Still the necessity of an appropriate initial guess in the NewtonRaphson method may prevent the solutions from converging [Jamet 73], and a
continuation technique, is often required.
The mixed formulation so-implemented is still quite expensive in terms of storing
memory requirements, with the associated coefficient matrix of the resulting system
being 2M+N dimensional, where M and N are the number of the velocity and pressure
unknowns respectively. So as to avoid this problem we are going to introduce the
penalty and the segregated formulations.
Penalty laminar formulation
The main difficulty found when obtaining a numerical solution for the Navier-Stokes
equations is that apart from verifying the dynamic constitutive equation, the solutions
must satisfy in addition the incompressibility condition. The mixed finite element
formulation leads to a system in which both velocity and pressure are taken as
unknowns. Nonetheless, besides the problems entailed in the election of the basic
elements in order to allow for the div-stability condition to be held, mixed methods
6
result in a large-dimensioned system. Therefore, not only a larger dimension has to be
handled with its corresponding increased memory requirements, but also the stiffness
matrix is found to be radically different to the narrow-band type of matrix, which is
preferred for the direct resolution of the system of equations.
To overcome these shortcomings, the penalty formulation is presented here. The
penalty formulation provides with the possibility of imposing the incompressibility
constraint without solving an auxiliary pressure equation, by replacing the continuity
equation with the expression:
(10)
ui ,i = −εp
where the so-called penalty parameter ε is a number close to zero. This equation is
incorporated into the dynamic equation, and therefore a system that depends on both
velocity and pressure is transformed into a velocity-dependant single equation, that
converges to the fully incompressible problem as ε approaches zero [Hughes 79],
[Heinrich 81], [Sohn 90], [Hannani 95]. Once we have applied the weighted residuals
method, the following integral dynamic equation is obtained:
1 h h
h
h
h h
h
h
h h
∫ wi ui ,t + u j ui , j − f i dΩ +ν ∫ wi , j ui , j dΩ + ∫ ui ,i wi ,i dΩ − ∫ ti wi dΓ2 +
Ωh
(
)
Ωh
ε
( )
Ωh
Γ2
1
h
(11)
+ ∑ ∫ pih uih,t + u hj ui,jh −νui,jj
− uih,i ,i − f i dΩ = 0
ε
e Ωe
which is solved for pressure. Once the velocity field has been obtained, the pressure
field can be calculated as a post-processing result, by using the formula:
1
(12)
p h = − uih,i
ε
The solution to equation (11) will approximate that of the initial problem as ε tends to
zero, provided that the penalty consistency condition is verified. If not, the use of the
penalty formulation could lead to the obtaining of a non-singular coefficient matrix
associated to the penalty term. As ε tends to zero, this term may dominate the system
of equations, therefore the whole problem could be over-constrained, and the only
possible solution could be the trivial one. When carrying out an exact integration of the
penalty term, ‘locking’ occurs and the only possible solution is the trivial one. This is a
problem totally analogous to the one obtained when a linear velocity and a constant
pressure is employed when using a mixed formulation. The discrete formulation in (11)
would not be consistent and the algorithm would not achieve convergence [Hughes 79].
This problem can be avoided by making a so-called selective reduced integration of the
elementary matrices involved in the resolution of the problem. A reduced numerical
integration consists in using a quadrature rule that is not exact for the polynomials
considered. The use of a one point Gauss quadrature rule for the integration of the
quadratic functions in the penalty term, transforms the associated ‘penalty’ matrix into a
rank deficient matrix and consequently ‘unlocks’ the obtaining of a non-trivial solution.
For more details on this topic you can refer to [Carey 84].
Once the elementary matrices are evaluated and assembled, the integral equation in (11)
can be expressed in matrix notation as:
Mυ
1
∂υ
+ Cυ (u, v )υ + νA υ υ + B ε υ = f
∂t
ε
7
(13)
where M υ is the mass matrix, Cυ (u, v ) is the convective matrix, A υ is the viscous
matrix, B ε is the penalty matrix, u is the velocity vector in the x direction, v is the
velocity vector in the y direction, f is external forces vector and υ is the velocity vector.
The non-linearities and the derivatives with respect to time are solved by using a
successive approximation method and a backward differencing scheme, in the same way
as we proceeded for the mixed formulation.
Segregated laminar formulation
To overcome the drawbacks arising from the resolution of the integrated velocitypressure and penalty formulations of the viscous flow, the so-called segregated methods
are developed in order to reduce the memory requirements when solving the NavierStokes equations. Some of the most commonly used of these segregated methods, that
obtain the flow variables in a sequential way, are the fractional step method [Donea 82],
[Ramaswamy 92], [Choi 97], and those based upon a SIMPLE algorithm [Benim 86],
[Rice 86], [Choi 94], [du Toit 98]. An algorithm based upon the SIMPLE method, first
released for finite volumes, is described in this section.
The penalty method succees in solving the Navier-Stokes Equations with great memory
savings due to the smaller number of equations to be solved, producing meaningful and
stable solutions thanks to the use of the so-called reduced integration as seen in the
previous section. Anyhow, the accurateness of the method depends on the election of
the parameter ε . For very small values of ε , the weight of the penalty term in the
stiffness matrix happens to cancel the amount of information contributed by the viscous
term, which is very small in comparison. This information is consequently truncated
and dropped from the equations. The penalty parameter should consequently be chosen
depending on the word length of the computer. On the other hand, if the selected
penalty parameter is too large, this choice may spoil the whole procedure, as ε is
wanted to tend to zero so as to allow for convergence. Consequently, the choice of ε is
not a trivial task, and a wrong selection in the parameter may lead to a meaningless
solution. Moreover, the penalty formulation achieves a great reduction in the storing
requirements, compared to the mixed formulation (the 2N+M equations in the mixed
formulation are reduced to a 2N dimensioned system in the penalty formulation). Still,
the stiffness matrix is far from being a narrow band type of matrix despite the
renumbering of the nodes.
Many of these shortcomings are not present in the so-called segregated methods, which
are broadly employed by many authors so as to solve the Navier-Stokes equations.
Following the success of the Finite Volumes Method [Patankar 80], several authors
adopted the formulation in the SIMPLE and SIMPLEST methods to the finite element
approach [Benim 86], [Rice 86], [Shaw 91], [Haroutunian 93], [Ferzinger 96]. These
segregated finite element schemes give solution to the problem of the viscous
incompressible flow, by employing a procedure in which the velocity and pressure
unknowns are not obtained simultaneously but in a sequential way. The segregated
formulations calculate velocities and pressures in an alternative iterative sequence,
requiring much less storing needs than the mixed methods. Moreover, these algorithms
not only achieve a greater reduction in the number of equations compared to the penalty
method (in this formulation the dimension of the system to be solved is equal to the
number of nodes), but also allow for the production of narrow band stiffness matrices,
when a proper renumbering of the nodes has been carried out. The segregated method
also avoids the use of the sometimes inconvenient penalty parameter.
8
Another gain of these segregated algorithms is that a mixed-order interpolation can be
used [Schneider 78], [Rice 86]. As has already been said, the mixed and penalty
methods require a velocity approximation different from that of the pressure. The
easier-to-implement discretization of the domain in terms of the same basic functions
for both velocity and pressure, leads to oscillation-free solutions, and the tendency to
produce the checkerboard pressure distribution is therefore eliminated.
The use of the weighted residuals method on the dynamic equation gives the dynamic
system:
h
h h
h
h
h
h h
h h
∫ wi u j ui , j − f i dΩ +ν ∫ wi , j ui , j dΩ + ∫ wi p,i dΩ − ∫ ti wi dΓ2 +
Ω
(
)
(
Ω
Ω
)
Γ2
h
+ ∑ ∫ pih u hj ui,jh −νui,jj
+ p,hi − f i dΩ = 0
e Ωe
(14)
The matrix version of the dynamic system can be written as:
∂N j
p j dΩ
C(u, v )u + νAu = Gu = f x − ∫ wi
Ω
∂x
∂N j
(15)
C(u, v )v + νAv = Gv = f y − ∫ wi
p j dΩ
Ω
∂y
where the right-hand side member of these equations is considered as a given value and
will be taken as zero for the first iteration. We can rewrite now equation (14) as:
∂N j
g ii ui = −∑ g ij u j + f xi − ∫ wi
p j dΩ
Ω
∂x
j ≠i
∂N j
(16)
g ii vi = −∑ g ij v j + f yi − ∫ wi
p j dΩ
Ω
∂y
j ≠i
where g ij is the coefficient matrix of the dynamic system. Now we can express the
velocities as:
∂N j
1
− ∑ g ij u j + f xi − ∫ wi
Ω
p
d
ui =
j
Ω
g ii j ≠i
∂x
∂N j
1
− ∑ g ij v j + f yi − ∫ wi
(17)
p j dΩ
vi =
Ω
g ii j ≠i
∂y
A set of ‘intermediate’ velocities, the pseudo-velocities u~i y v~i , are defined as follows:
1
1
− ∑ gij u j + f xi
− ∑ gij v j + f yi
u~i =
v~i =
(18)
gii i ≠ j
gii i ≠ j
The velocities can be now approximated in terms of the pseudo-velocities by using the
expressions:
∂N j
∂N j
(19)
pj ;
ui ≈ u~i − K ip
vi ≈ v~i − K ip
pj
∂x
∂y
Where the velocity-pressure coupling coefficients K ip are given by the equation:
1
(20)
K ip =
wi dΩ
g ii ∫Ω
Using the weighted residuals method in the continuity equation we obtain:
h
h
h h h
(21)
∫ wi , j u j dΩ − ∫ wi u j n j dΓ2 = 0
Ωh
Γ2
or in expanded form:
9
∂N j
∂N j
∂wi
∂wi
p
p
N
K
p
+
N
K
p j dΩ =
k
k
j
k
k
∫Ωh ∂x
∂x
∂y
∂y
∂wi ~ ∂wi
(22)
=∫
N ju j +
N j v~j dΩ − ∫ wi (N j u j n x + N j v j n y ) dΓ2
Ω h ∂x
Γh
∂y
Once the velocity system has been calculated, the pressure system is obtained and
solved, the velocities are then corrected by using the following expression:
∂N j
∂N j
1
~ − 1 w
p
d
Ω
p j dΩ
u i = u~i −
w
=
v
v
(22)
i
j
i
i
i
∂x
∂y
g ii ∫Ω
g ii ∫Ω
to ensure continuity.
The iterative process is based upon assuming a zero pressure field as a first guess in the
resolution of the dynamic equation, providing the velocity field as the output. Once the
pseudo-velocities and the pressure-velocity coupling coefficients have been calculated,
the continuity system is assembled and solved, and thus the values for the pressure field
are obtained. Finally, the velocities are updated, making use of the newly determined
pressure field, and with both the new velocities and pressures the dynamic equations are
reassembled, solved and the same procedure is repeated until convergence is achieved.
When using a segregated algorithm, the use of uncoupled velocity and pressure fields
may lead to the divergence of the whole process. To avoid this problem, an underrelaxation of the unknowns can be introduced so as to guarantee the convergence of the
process. The linear relaxation formulae to be used for this purpose is:
(23)
φ n = φ n −1 + α φ n − φ n −1
where φ n and φ n−1 are the values of the unknowns (either velocity or pressure) at the
present and former iterations. The momentum equations are also under-relaxed making
use of an inertial relaxation factor r i defined as:
(
)
ri = ∫ wi dΩ
(24)
Ω
with r i being added to the terms in the diagonal of the dynamic coefficient matrix as
follows:
∂N
(g ii + ri )uin + ∑ g ij u nj = f xi − ∫Ω wi j p j dΩ + ri uin−1
∂x
j ≠i
∂N
(g ii + ri )vin + ∑ g ij v nj = f yi − ∫Ω wi j p j dΩ + ri vin−1
(25)
∂y
j ≠i
with uin−1 and vin−1 being the values of the velocities obtained in the previous iteration.
The use of this formulation, leads to a N-dimensioned narrow band coefficient matrix,
and consequently to further memory savings in the resolution of the Navier-Stokes
equations.
Shallow Water formulation
For the resolution of the Shallow Water equations we will use now a Finite Element
mixed approach that will share the same advantages and shortcomings of that of the
laminar mixed formulation explained before. If we apply the weighted residuals method
on both the dynamic and continuity Shallow Water equations, we obtain:
h
h
h h
h
h
h
h
h h
h h
∫ wi ui,t + u j ui, j − g S0i − S fi dΩ + ν ∫ wi, jui, j dΩ − g ∫ wi,i h dΩ − ∫ h ti wi dΓ2
Ωh
(
(
))
Ωh
10
Ωh
Γ2
+∑
∫ p (u
e Ωe
h
i
h
i ,t
(
))
+ u hjuih, j − νuih, jj + gh,hi − g S0hi − S hfi dΩ = 0
∫ q (h
h
Ωh
h
,t
)
+ h h uih,i + uih h,hi dΩ = 0
(26)
But now we have a new difficulty that did not appear in the numerical approach to the
2D Navier-Stokes equations presented previously: we have the depth itself and the
gradient of depth being included as part of the continuity equation. In fact, the inclusion
of the depth and the gradient of depth in the continuity equation, allows for the
verification of the conservation of mass in a pseudo-3D basis and not on a 2D laminar
sense, as a consequence of having carried out an integration in depth of the NavierStokes equations. As a consequence, some pseudo-non-linearities show up in the
continuity equation, which should be considered in addition to the non-linearities
resulting from the convective quadratic term. The Shallow Water equations will be
integrated in order to cope with this problem.
Let us introduce the following approach: we are going to assume that the depth values
in the continuity equation are going to be constant all over the domain for the first
iteration, and equal to the outflow given depth. In the following iterations carried out in
order to solve the convection, the depths and gradients of depth in the continuity
equation will be evaluated from the results of the former iteration, and this evaluation
will be carried out in terms of a finite difference approach. Since a non-equal order
interpolation of the unknowns must be used to achieve converge, the velocities and the
depths are calculated on a different mesh. The depths to be re-fed in the continuity
equation for the second and the following iterations are going to be evaluated on the
velocity mesh points. Recall that the basic element used in this formulation is the Q1P0
basic element, or in other words, the velocity is interpolated in terms of bilinear
continuous functions with respect to a four-nodded basic element, and the depth is
interpolated in terms of constant discontinuous functions within the basic element. The
depth at a velocity node hi* will be taken as the mean value of the depths for the former
iteration in the surrounding basic elements, and the gradients of depth on the velocity
mesh h * , will be evaluated from the star depths hi* on a finite difference basis, i.e.:
hi* =
1 n k
∑ hi
n k =1
hi*, x =
hi*, y
1 n
∑
n k =1
1 n
= ∑
n k =1
h*1i1
∑x
3
j =1
*
− hi*
hkij
∑y
3
j =1
− xi
h − hi*
kij
*
kij
kij
− yi
h*1i2
k=1
k=4
h*i
(27)
k=3
h*1i3
k=2
where n takes a value of 1, 2, 3 or 4 depending on the velocity node being a convex
corner, a side, a concave corner or an inside node, and h i j is the constant depth in the
*
are the star depths on the velocity nodes in the basic
surrounding elements. Where hkij
element k, that shares a common node (i) with hi* , xkij and y kij are the x and y coordinates of the nodes in the basic element k, that shares a common node (i) with hi* , xi
and yi are the co-ordinates of the node where the gradient of depth is being evaluated
and n is defined in the same way as for the star depths. The contribution to the
11
derivative with respect to x by nodes with the same abscise is ignored, and so is the
contribution to the derivative with respect to y by nodes with the same ordinate, in order
to avoid a division by zero.
After each iteration for convection has been solved, the star depths and star gradients of
the depth are calculated and re-fed into the continuity equation. The use of this
algorithm developed by the authors in the resolution of the Shallow Water equations
achieves very good numerical results as will be seen in the numerical examples. The
general procedure for the obtaining of the steady system of differential equations could
be written in its matrix form as:
Cυ (u, v )υ + νA υ υ − Bh = f
(28)
D h * υ + E h * υ = 0
( )
( )
( )
where C υ (u, v) is the convective matrix, A υ is the viscous matrix, B is the depth matrix,
f is external forces vector, D(h*) is the star depth matrix, E h * is the star gradient of
depth matrix, f is the external forces vector, h is the depth vector and υ is the velocity
vector.
SUPG stabilization of the algorithms
The above formulations include a SUPG term in the dynamic weighting functions. This
SUPG contribution will affect all the terms in the dynamic equation and will be based
upon an optimal rule function for the obtaining of the multi-dimensional diffusion
coefficient. The SUPG formulation will be the following:
k uˆ hj wih, j
with
(29)
pih =
wi = wi + pi
uh
where the multi-dimensional definition of the diffusion coefficient k is given by:
ξ uξh hξ + η uηh hη
u
2
uˆ i = i ,
u = ui ui
k =
u
2
where
1
1
ξ = cothα ξ −
η = cothαη −
αξ
αη
u hh
u hh
αξ = ξ ξ
αη = η η
2ν
2ν
uξh = eξi ueih
uηh = eηi ueih
η
(30)
ξ
1
hη
eη
eξ
ξ
-15
x2
-10
-5
5
-1
x1
hξ
Figure 1. Characteristic basic-element lengths and unit vectors.
Optimal rule for the approximation of ξ y η .
12
10
15 α
where hξ , hη and eξi , eηi are the characteristic basic-element lengths and unit
vectors in the direction of the local axes ξ and η (see figure 1). The parameters α ξ
and αη are the directional Reynolds numbers of the basic element, u eih is the velocity
in the interior of the element and ν is the kinematic viscosity of the fluid. Different
versions of the streamline upwind formulation have been used by other authors and
can be found in [Franca 92], [Sampaio 91], [Hill 95], [Hannani 95], [Cruchaga 96],
and [Choi 97]. For the present work, the stabilization technique will be based upon
the streamline upwind Petrov-Galerkin weighting functions as defined in (29, 30).
These weighting functions will be applied on the formulation as specified in
formulas (5, 11, 14, 26), with very good results as will be seen in the numerical
examples shown in the following section.
NUMERICAL RESULTS
Once we have presented the algorithms to be used in this work, we are going to proceed
to validate them by using some benchmark problems with known solution.
Validation of the laminar algorithms, the Cavity flow and the Backward-facing
step benchmark problems
The Cavity flow benchmark problem
The driven cavity flow is a classical test used by many authors to check the quality of
the methodology employed in the resolution of the 2D Navier-Stokes equations. This
benchmark problem is based upon the analysis of the flow in a square cavity with
prescribed horizontal velocity in the upper side and solid boundaries in the lateral and
bottom sides. This is a challenging problem due to the presence of several re-circulating
regions in which the solution changes rapidly, and because of the pressure singularities
that show up in the upper corners. This benchmark test will be used, to validate the
algorithms developed in this thesis by its comparison with the results obtained by other
authors. These results to compare with, will be those of Ghia [Ghia 82] obtained by
employing a mesh of 129x129 nodes; Hannani [Hannani 95] with non-uniform meshes
of 32x32, 45x45 and 80x80 Q1P0 basic elements; and the results by Kondo [Kondo 91],
making use of a 40x40 element mesh of four-node, non regular basic elements. All of
them can be considered as reference results, specially those of Ghia, that are commonly
employed to check the validity of the algorithms by most of the authors in the related
bibliography.
The boundary conditions used for this problem have been of the Dirichlet type in all the
boundaries. A unitary horizontal velocity heading towards the right hand side has been
prescribed for the top side (including the upper corners), and the no-slip condition has
been considered for the rest of the boundary. The discretized domain used in the
calculations has been a 1681-node non-regular mesh with 1600 Q1P0 elements (see
figure 2).
13
400
350
300
Y
250
200
150
100
50
0
0
100
200
300
400
X
Figure 2. Cavity Flow 41x41 non-regular mesh
The results for all the formulations considered have been the same and can be seen in
figures 3-6. This benchmark problem has been solved by making used of the mixed,
penalty and segregated algorithms and the Reynolds numbers used have been 100,
1000, 5000 and 10000. The Reynolds number has been defined as Re = U·L / ν, where
U is the velocity in the upper side, L is the length of the side of the square domain, and
ν is the kinematic viscosity. The value of Reynolds = 10000 is considered as a limit for
the steady Cavity Flow calculations, since has been shown through detailed numerical
experiments [Shen 76] that above this bound, the stationary solution ceases to be stable.
400.00
400
400
350
350
300
300
250
250
200
200
Y
Y
350.00
150
150
100
100
50
50
vel
0.9375
0.875
0.8125
0.75
0.6875
0.625
0.5625
0.5
0.4375
0.375
0.3125
0.25
0.1875
0.125
0.0625
300.00
250.00
200.00
150.00
100.00
50.00
0
0
100
200
300
400
0
500
0
100
200
300
400
500
X
X
0.00
0.00
50.00
100.00
150.00
200.00
250.00
300.00
350.00
400.00
250.00
300.00
350.00
400.00
250.00
300.00
350.00
400.00
Figure 3. Cavity flow. Velocity field, streamlines, and pressure field Re=100
400.00
400
400
350
350
300
300
250
250
200
200
Y
Y
350.00
150
150
100
100
50
50
vel
0.9375
0.875
0.8125
0.75
0.6875
0.625
0.5625
0.5
0.4375
0.375
0.3125
0.25
0.1875
0.125
0.0625
300.00
250.00
200.00
150.00
100.00
50.00
0
0
100
200
300
400
0
500
X
0
100
200
300
400
500
X
0.00
0.00
50.00
100.00
150.00
200.00
Figure 4. Cavity flow. Velocity field, streamlines, and pressure field Re=1000
400.00
400
400
350
350
300
300
250
250
200
Y
Y
350.00
vel
0.9375
0.875
0.8125
0.75
0.6875
0.625
0.5625
0.5
0.4375
0.375
0.3125
0.25
0.1875
0.125
0.0625
200
150
150
100
100
50
50
300.00
250.00
200.00
150.00
100.00
50.00
0
0
100
200
300
400
500
0
0
100
200
300
X
X
400
500
0.00
0.00
50.00
100.00
150.00
200.00
Figure 5. Cavity flow. Velocity field, streamlines, and pressure field Re=5000
14
400.00
400
400
350
350
300
300
250
250
Y
Y
350.00
200
Level
19
18
17
16
15
14
13
12
11
10
9
8
7
6
5
4
3
2
1
200
150
150
100
100
50
50
vel
0.95
0.9
0.85
0.8
0.75
0.7
0.65
0.6
0.55
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
300.00
250.00
200.00
150.00
100.00
50.00
0
0
50
100
150
200
250
300
350
400
450
0
500
0
50
100
150
200
250
300
350
400
450
500
0.00
0.00
X
X
50.00
100.00
150.00
200.00
250.00
300.00
350.00
400.00
Figure 6. Cavity flow. Velocity field, streamlines, and pressure field Re=10000
The most commonly used comparison results for this benchmark problem are the
horizontal velocities along a vertical central line, which have been plotted in figure 7.
Re=100
Re = 1000
Re=100
1.2
1.2
1.2
1.0
0.8
1.0
Present 41x41
Ghia 129x129
0.6
1
0.4
0.2
0.8
0.8
0.0
-0.5
0
0.5
1
1.5
0.6
Ghia 129x129
Present 41x41
0.4
0.4
Hannani 45x45
0.2
0.2
Present 41x41
Ghia 129x129
0.6
0.0
-0.5
Hannani 80x80
0
0
0.5
1
-0.5
1.5
0
0.5
1
1.5
Figure 7. Horizontal velocities along a central vertical line for Re=100, and 1000.
Re = 10000
Re = 5000
1.2
1.2
1
1
0.8
0.8
Ghia 129x129
Present 41x41
0.6
0.4
0.4
0.2
0.2
0
-0.5
Ghia 129x129
Present 41x41
Kondo 40x40
0.6
Hannani 31x31
0
0.5
1
0
-0.5
1.5
0
0.5
1
1.5
Figure 8. Horizontal velocities along a central vertical line for Re= 5000, and 10000.
The horizontal velocity profiles along a central vertical line adjust to the reference
values of [Ghia 82], with a much finer mesh and are also substantially better than those
of [Hannani 95] and [Kondo 91], for a mesh with a similar refinement and even a finer
one. No substantial differences are observed among the results of the three formulations
used for the velocity results nor for the pressure field results, which are also in good
agreement with the benchmark solutions of the problem obtained by those authors.
The good results obtained in the velocity profiles have made useless the employment of
a finer mesh, which would necessitate a much longer CPU time. The calculation times
are shorter for the mixed algorithm and of increasing magnitude for the penalty and
segregated method. For the penalty solution, the introduction of the penalty parameter
makes the linear system of equations more difficult to be solve, since the penalty
parameter tends to zero. This computational time can be reduced, anyhow, by the use of
a properly weighted penalty parameter. For the segregated resolution of the flow, a
direct solver has been used in the calculations, with a definitively greater computational
cost, and the convergence process is consequently slowed down.
The algorithms implemented have proved to give very accurate results even for a less
refined mesh, showing that the upwind weighting implemented in the numerical scheme
15
is a powerful tool to solve some flow problems without using very refined meshes, and
with no wiggles in the so-obtained solution.
The Backward-facing step benchmark problem
The laminar Backward Facing Step benchmark problem is presented next, as one of the
most commonly used benchmark problems in the literature, in order to validate the
algorithms that give solution to the Navier-Stokes equations. The backward step is
based upon a simple geometry where flow separation and reattachment occur.
Experimental data for this problem can be found in [Armaly 83], who also solved this
problem numerically by using a control-volume-based Finite Difference Method. The
problem of the backward step flow will be solved in this section by using the penalty
algorithm and its results will be compared with those of Armaly, which are generally
used as verification data.
The geometry and boundary conditions considered for this benchmark problem have
been those used in [Armaly 83]. An expansion ratio of 1:1.94 has been considered for
the widening of the channel, which has a total length of 50 so as to allow for the
vortices to take place. The inlet boundary has been located at 3.5 step heights upstream
of the expansion corner. The domain has been split into 2850 Q1P0 basic non-regular
elements with 3021 nodes. The mesh is coarser at the outlet and more refined at the lefthand side of the channel, so as to allow for a better accuracy in the regions where the
primary vortices occur. A bias parameter of 0.5 has been used for this purpose along the
x-axis, therefore the width of the basic elements at the inlet is one half of that of the
elements at the outlet, and the height of the basic elements is uniform within the whole
domain. The mesh can be seen in figure 9, where a magnifying factor of two has been
used for the y-axis. A parabolic horizontal velocity profile has been imposed at the inlet
with a maximum velocity of 1, and the velocity is equal to zero at the boundaries. The
lateral sides have been considered as solid boundaries and the no-slip condition has
been imposed on them. Finally, a zero traction condition has been imposed at the outlet.
4
3
2
1
0
0
5
10
15
20
25
30
35
40
45
50
X
Figure 9. Backward Facing Step. Mesh
The flow has been obtained for a Reynolds number between 100 and 1200. The
Reynolds number has been defined as Re = u·D / ν, where u is the average inlet
velocity, D is the hydraulic diameter and the kinematic viscosity ν has been altered so
as to make the Reynolds number vary. As foretold by the experimental results in
[Armaly 83], there exists a single re-circulation zone at the expansion corner up to a
Reynolds number of about 450, beyond which a second vortex shows up at the top
boundary, and gets bigger as the Reynolds number is increased.
2
Y
1
0
0
5
10
15
X
Figure 10. Flow in a Backward Facing Step. Streamlines for Reynolds = 400
16
3
Y
2
1
0
0
2
4
6
8
10
12
14
16
18
20
X
Figure 11. Flow in a Backward Facing Step. Streamlines for Reynolds = 500
8
7
6
Y
5
4
3
2
1
0
0
5
10
15
20
25
30
35
40
45
50
X
Figure 12. Flow in a Backward Facing Step. Streamlines for Reynolds = 1200
The size of the reattachment zones si versus the Reynolds number is compared with the
experimental and numerical results by Armaly; these results can be seen in figures 10,
11 and 12. The reattachment locations of the vortices are defined as follows; s1 is the
reattachment location of the primary vortex, s2 is the separation location of the
secondary top boundary vortex and s3 is the reattachment location of the secondary
vortex. All of them have been measured from the expansion corner, as depicted in figure
13.
s3
s2
s1
Figure 13. Flow over a Backward Facing Step. Sketch of the recirculation lengths
As seen in figures 14, 15 and 16, the computed results obtained in the present work
compare more favourably with experimental data than the numerical results from
Armaly. Although the present results are totally analogous to the experimental data in
[Armaly 83] for s3 and for all the Reynolds numbers considered, when taking about s2
and specially s1, the experimental data differ from the calculated results beyond a
Reynolds number of about 400. This difference between measured and calculated values
is not only shown in the numerical results by Armaly, but also in the results by [Kim
88] and [Kwack 85] among many others. The differences in these values are due to the
fact that the 3D effect becomes very important as the Reynolds number is increased. As
pointed out by Armaly, these effects became predominant beyond a Reynolds number
of 1300.
Reattachment length s1
20
18
16
14
12
10
8
6
4
2
0
s1 Armaly exp
s1 Armaly cal
Present
0
200
400
600
800
1000
1200
1400
Figure 14. Reattachment length s1 versus Reynolds number for the Backward Facing Step
17
Reattachment length s2
18
16
14
12
s2 Armaly exp
10
s2 Armaly cal
8
Present
6
4
2
0
0
200
400
600
800
1000
1200
1400
Figure 15. Reattachment length s2 versus Reynolds number for the Backward Facing Step
Reattachment length s3
30
25
20
s3 Armaly exp
s3 Armaly cal
15
Present
10
5
0
0
200
400
600
800
1000
1200
1400
Figure 16. Reattachment length s3 versus Reynolds number for the Backward Facing Step
Verification of the Shallow Water Algorithm
The Shallow Water algorithm has been verified by making use of the numerical
example of a channel in which the width is sharply doubled in the flow direction. The
flow in the widening channel has been obtained by using the mixed Shallow Water
equations with a kinematic viscosity equal to 10-6 m2/s. The channel has a length of 100
m, and spreads from 10 meters of width at the inlet, up to 20 m at a distance of 13 m
from the inlet. A hydrostatic pressure of 1 m has been imposed at the outlet, and a
roughness Manning coefficient of 0.01 m–1/3s has been considered throughout the
channel length. The domain has been split into a 1171-node, non-regular mesh with
1090 Q1P0 basic elements (see figure 17).
Frame 004 22 Jun 2000 ITERAC=
2Visc= 1.000000000000000E-006
50
Y
40
30
20
10
0
0
25
50
75
100
X
Figure 17. Widening canal. Finite Elements Mesh
An inflow normal uniform velocity of 3 cm/s has been imposed along the inlet and an
adverse slope against the main flow direction of magnitude 10-2 has been considered all
over the domain. For these flow conditions, a parallel flow is achieved at the expansion
corner. As we are moving within a Froude number much smaller than the unity in all
these examples, the flow may be described as subcritical with Fr<<1, where the Froude
number is defined as:
18
Fr =
v
gh
(31)
where v stands for the one-dimensional velocity, g is the gravity acceleration and h is
the depth. The general equation of the one dimensional gradually varied flow (see for
instance [Chadwick 86]) can be written as:
dh S 0 − S f
=
dx 1 − Fr 2
(32)
where dh/dx is the variation in depth along the length of the channel, and So and Sf are
the geometric and friction slopes respectively. For the present conditions, the formula
32 can be simplified into:
∆x 1
≈
∆h S 0
(33)
The x-component of the velocity is plotted along the domain for both the twodimensional Navier-Stokes and the Shallow-Water algorithms, so as to compare them.
As can be seen in plot 18, the Shallow Water algorithm behaves following (33). For the
Navier-Stokes formulation the discharge at the inflow is 0.440 m3/s, and at the outlet the
discharge is 0.286 m3/s. The continuity equation is not verified as this algorithm can
only be applied to the simplification of a 2D flow. On the contrary, when the ShallowWater equation is used both discharges at the inlet and at the outlet are equal to 0.587
m3/s. The mass is therefore conserved along the channel. As can be seen from these
examples, the 2D laminar Navier-Stokes equations can provide a good evaluation of the
shallow flow in which no major changes in the depth are taking place, but when an
important variation in the depth occurs, the Shallow Water formulation is a more
appropriate formulation that allows for the conservation of mass throughout the domain.
Frame 001 25 Jun 2000 ITERAC=
Frame 001 25 Jun 2000 ITERAC=
6Visc= 1.000000000000000E-006
6Visc= 1.000000000000000E-006
Z
Z
Y = 3 cm/s, S = -10-2). Velocity module and depths
Figure 18. Flow in a widening canal (vel
0
X
Laminar Navier-Stokes (lower red surfaces)
versus Shallow Water(upper blue surfaces)
Y
X
2
1.75
0.04
1.5
VELX
1
0.02
0.5
20
0.01
20
0.75
0
0
0.25
40
40
X
0
20
60
80
10
100
0
X
60
0
20
80
Y
10
100
19
h
1.25
0.03
0
Y
APPLICATION TO SOME WASTEWATER TREATMENT PROBLEMS
Once the code has been checked on some well-known benchmark problems with
optimum results, it has been used to solve some real flow problems related with the civil
engineering technology and in particular with the wastewater treatment industry.
Some authors [Espert 96] have used the potential flow equations to evaluate the flow in
clarifiers and other wastewater treatment basins. When we use these simplifications, we
can obtain an approximation of the flow for slow creeping conditions, but only the
resolution of the all-term-including Navier-Stokes equations will allow us to detect the
real streamlines and the vortices that show up even for very slow water flows. Let us
now use the previously explained formulations in the resolution of some wastewater
treatment basins.
Flow in clarification basins
The flow of water in a rectangular and circular conventional clarification basins has
been considered. Clarification has two main applications in the water treatment
processes. Its most usual aim is to reduce the solids load after coagulation and
flocculation have taken place. Its second application is the removal of heavy settleable
solids from a turbid source to lessen the solids load in water. The aim of a good
clarification basin design is the obtaining of a sufficiently stable flow, so as to achieve a
better sedimentation. There is a large number of non-conventional devices for high rate
clarification, such as tube or plate settlers, dissolved air flotation clarifiers, sludge
blanket or slurry recirculation clarifiers. The choice of one of those depends on the
features of the inflow water, the outflow water requirements, time, space and budget
availability to carry out the purification of the water. The flow in a prototype of plate
settlers with biological treatment, being developed in the School of Civil Engineering of
La Coruña (ETSICCPC), has been also regarded. The description of the flow may be a
powerful tool to attain an optimum shape in the designing of these structures, in order to
make the most of the plant resources.
Flow in a rectangular and circular conventional clarifiers
The rectangular and circular basins are the most commonly used clarification devices.
In spite of their simplicity, they have achieved excellent results with scant maintenance
costs. These basins were originally designed with the capacity to store sludge for
several months and were periodically taken out of service for manual cleaning. Today,
most of the clarification basins include a continuous cleaning mechanical equipment,
such as dragging chains that plow the sludge along the basin floor to hoppers.
Nevertheless, these mobile devices for cleaning and other purposes do not have an
important influence in the streamline distribution, and can be ignored when the flow is
calculated (for further details on clarification basins you can refer to [Metcalf 95]). We
are going to evaluate the flow features in the vertical section of a rectangular and
circular clarifiers, with the following dimensions and design parameters:
Clarifier
Rectangular
Circular
Dimensions (m)
Depth (m) Slope (%)
Nodes&Elements
Detention time (h)
Surf. Load. Rate (m/h)
W: 9, L: 24
3.3
1.2
1052, 949
3
1
D: 17.5
3.65
0.8
817x2, 756x2
3
1
20
Walkway
Influe
inta
Influent
intake
Foam launder
Baffle plate
Foam sweeper
Foam sweepers
Overflow
launder
Baffle
plate
Foam
launder
Baffle slab
Overflow
launder
Effluent
pipe
Sluddge scrappers
Sluge hoppers
Influent
pipe
Sluge scraper
Sludge removal pipe
Sludge hopper
Sluge withdrawal
Figure 19. Rectangular and circular clarifier with bottom sludge scraper. Sketch
When working with clarifying basins, one of the criteria to be used in their definition is
that of achieving a maximum head loss at the inlet, so as not to disturb the slow flow of
the water mass being treated. Therefore, we should avoid turbulence by placing some
kind of energy dissipating structure in the faster zone, that is the inlet (see figure 19).
One of these maze-looking dissipating structures has been considered for the inlet of our
clarifiers, being placed in the left-hand side for the rectangular one and in the centre of
the basin for the circular, where the respective inlets are. For the outlets (in the righthand side for the rectangular one and in the perimeter for the circular one), a
conventional overflow launder has been disposed. The domain in which the flow takes
place has been split into 949 Q1P0 basic elements with 1052 nodes for the rectangular
and 817x2 nodes and 756x2 elements for the circular, which has benefited from its
symmetry property. For the working parameters chosen, a velocity of 1 cm/s has been
imposed at the inlet in both of them. The no-slip condition has been imposed at the
bottom and lateral sides, and the spillway has been left free with a zero traction
boundary condition. For the topside, the vertical velocity has been fixed as zero and the
horizontal velocity has been left free. The results for the velocity and pressure fields can
be seen in figures 20 and 21.
Frame 001 29 Jun 2000 ITERAC
Frame 001 01 Aug 2000 ITERAC
700
600
800
700
500
Y
600
Y
500
400
400
300
300
200
200
100
0
100
0
1000
2000
X
0
-800
-700
-600
-500
-400
-300
-200
-100
0
100
200
300
400
500
600
700
800
900
X
Figure 20. Flow in a rectangular and circular clarifying basin. Streamlines
300.00
200.00
200.00
100.00
200.00
600.00
1000.00
1400.00
1800.00
2200.00
100.00
200.00
300.00
400.00
500.00
Figure 21. Flow in a rectangular and circular clarifying basin. Pressures
21
600.00
700.00
800.00
Flow in a lamellar ‘LUPA’ clarifier
The behaviour of the water flow in a lamellar settler being developed in the ETSICCPC
is also regarded. In the plate settler, the settling area is increased by the disposal of
several lamella plates along the basin. By doing so, the settled solids in the plates are
dropped and removed, and therefore the settling rate is improved. Figure 22 shows the
cross-section of a standard plate settler. The environmental and sanitary engineering
area of the ETSICCPC is developing the design of a lamellar settler equipped with a
permeable bio-film that carries out a biological treatment of the water. As part of this
researching work, a prototype of the ‘LUPA’ clarifier (see Picture 1) has been
constructed and the flow along the model has been evaluated.
Discharge flumes
Lamella plates
Flocculation chamber
Sludge hopper
Figure 22. Conventional plate settler. Cross section Picture 1. The ‘LUPA’ prototype
The code has been used to evaluate the flow in the interlamellar vertical section, with
dimensions 80x30x10 cm3, with a water inflow at the bottom, and a free overflow at the
top. The lamellas are placed at a 50.1 degree angle with respect to the horizontal. The
flow has been calculated with the same assumptions to those considered for the
horizontal flow clarifiers, that is, with no-slip condition on solid boundaries and vertical
velocity equal to zero on the free surface. The definition of the boundary conditions is
completed with a Dirichlet relationship at the inlet and a Newman equality with zero
traction at the outlet. The longitudinal section of the clarifier has been divided into 1760
Q1P0 basic elements and 1916 nodes. The velocity at the inflow is set parallel to the
walls. The flow has been calculated for a discharge of 15 l/day and 1500 l/day
corresponding with inflow velocities of 5·10-3 and 5·10-1 cm/sg.
Frame 001 30 Jun 2000 ITERAC=
1Visc= 1.000000000000000E-002
50
50
40
40
30
30
Y
Y
Frame 003 03 Jul 2000 ITERAC=
20
20
10
10
0
0
10
20
30
40
50
60
0
70
X
0
1Visc= 1.000000000000000E-002
10
20
30
40
X
Figure 23. Flow in the lupa prototype, q = 15 and 1500 l/day. Streamlines
22
50
60
From the plots, it can observe how two primary vortices show up at both sides of the
inlet for very slow flows (15 l/day). The bigger one is generated at the right-hand side,
which is the lower and larger side, and a smaller one can be seen in the left-hand side.
As the discharge is increased, not only the size of the primary vortices is increased, in
particular for the right-hand side one, but a secondary vortex showing up, can also be
seen. This secondary vortex not only alters the course of the flow, but also allows for an
increase in the contact time of the water with the upper lamella, and consequently with
the bio-film on it, improving the behaviour of the clarifier.
Flow in a maze flocculator
The flow along a maze chamber, often used in the flocculation processes, has been
calculated. Flocculation is defined as the agglomeration of small particles and colloids
to form settleable or filterable particles. A separate flocculation process, where chemical
aids are added to water, is very often included in the treatment train to enhance contact
of destabilized particles and to build dense floc particles of optimum size. The hydraulic
flocculators, in opposition to the mechanical ones, allow for the formation of the flocs
without the help of any mechanical device. This type of flocculation is simple and
effective, especially for relatively constant flows. This sort of chambers is also used in
chlorination processes. Chlorination forms part of the chemical disinfection treatments
that are carried out on supply water in order to achieve its purification and
transformation into drinkable water. The aim of this winding design is to achieve a slow
and steady flow over a long distance to allow for the flocs to form. In chlorine
disinfection processes, this slowness enables water to maintain contact with the
chemical reagent over a long period of time (see [Metcalf 95] for further details on
maze flocculators). The velocities involved are quite slow, and a laminar flow is
expected, however, small vortices can show up and the Stokes evaluation of the flow
could not detect them. For this reason, a convective-term-including formulation is
required.
A rectangular chamber, in which water is re-circulated along a winding path, often
constitutes this kind of basins, and for this particular case will be modelled as a
prismatic tank with dimensions 8m wide and 10 m long, in which a twisting channel is
inscribed, split into 10 straight segments. The design parameters chosen for the
chlorination tank are the following:
-
Tank dimensions: 8x10x2 m3
Channel width: 1m
Channel length: 80 m
Horizontal velocity: 6.6 cm/s
Contact time: 20.2 minutes
A 2091-node regular mesh with 2000 Q1P0 basic elements has been chosen to model
the tank. The mixed Shallow Water algorithm has been used with a Manning coefficient
of 0.012 m-1/3/s. A Dirichlet boundary condition has been prescribed at the inlet, where a
parabolic velocity of 6.6 cm/s has been settled at the six lower left-hand-side nodes. At
the outlet, the velocity on the six lower right-hand-side nodes has been considered as an
unknown, and a hydrostatic pressure boundary condition of 2 m depth has been
prescribed. A slope of 10-3 has been considered falling rightward all over the domain. A
viscosity of 1.0·10-6 m2/s has been used for the wastewater.
23
Frame 002 21 Jun 2000 ITERAC
800
700
600
Y
500
400
300
200
100
0
0
250
500
750
1000
X
Figure 24. Flow in a maze flocculator. Mesh
As can be seen in figure 25, the results for the velocity field show the appearance of
small re-circulation zones besides the corners. This re-circulation is the responsible for
both the appearance of sediments in this region and also is the cause of a certain energy
loss, as can be seen in the pressure plot (figure 25). These effects, if unwanted, could be
removed by decreasing the velocity of the flow or the re-shaping of the channel.
Frame 002 21 Jun 2000 ITERAC
Frame 001 22 Jun 2000 ITERAC
Z
900
X
Y
800
700
2.00004
500
2.00003
Y
600
2.00002
400
2.00001
300
0
2
100
200
0
200
300
100
400
200
500
300
100
600
400
Y
0
0
250
500
750
700
500
800
600
1000
700
X
X
800
900
1000
Figure 25. Maze flocculator. Streamlines and depths (in m)
CONCLUSIONS
In this work, an exhaustive analysis of the incompressible flow has been carried out,
from the very definition of the governing equations, up to the resolution of some
practical problems, passing through the comprehensive study of the numerical
techniques used in their resolution. As a direct consequence of this study, a code has
been written based upon this analysis, which allows for a modelling of the
incompressible flow based upon a realistic interpretation of the forces taking place
within the flow, and giving optimum results.
The three different approaches: mixed, penalty and segregated, have been implemented
and their results have been checked and verified by the comparison of the three of them
among themselves and also against some reference results. As a consequence, several
conclusions have been reached. The first is that, as expected, the results obtained by the
three of them in the resolution of some benchmark problems have been identical, in a
comparison study that had not been carried out prior to this work. The different
approaches result in a different computer efficiency that depends not only on the
algorithm employed, but also on the numerical solver used to obtain the solution to the
resulting system of equations. Nonetheless the algorithm used does not affect the
24
accuracy of the solutions when an adequate selection of the numerical parameters has
been carried out. The second conclusion is that all the results compare very favourably
with the reference numerical and empirical results by other authors. As a consequence,
the code not only enables a comparison study of the available finite element numerical
techniques for the resolution of the Navier-Stokes equations, but also, as proved by the
examples provided, contributes to a better and faster approach to these problems.
A newly developed algorithm for the resolution of the Shallow Water equation, making
use of the finite difference tools within the finite element frame, has been implemented
with optimum results. The evaluation of this friction term is based upon on a Manning
type formula, which makes use of the empirically determined Manning roughness
coefficient. This term accounts not only for the energy losses that take place because of
the friction with the wetted perimeter, but also for the overall turbulent losses that take
place over the whole domain of integration. The turbulent eddies taking place within the
flow are not detected, but the turbulent energy losses are taken into account thanks to
this empirically determined formula, which provides a meaningful solution for practical
flows.
Some of the most commonly used hydrodynamic models for the flow calculations (such
as the RMA2 model developed by the Brigham University, which is broadly used
world-wide), incorporate a turbulence model featured by a constant eddy viscosity,
which is not hydraulically speaking well justified. In contrast, the Shallow Water
algorithm developed by the author includes an empirically determined turbulent losses
term but also keeps the Navier-Stokes formulation of the problem, being ready to
incorporate a k-ε turbulent model that has been developed within the research group and
provides an eddy viscosity that varies in time and space.
The accuracy of the numerical solutions so-obtained has been checked by using some
reference benchmark numerical and empirical solutions with great success, and once the
program has been validated, it has been used in the resolution of some wastewater
treatment flow problems. The so defined code creates an optimum frame for the
evaluation of the flow in some wastewater treatment basins, which is an essential tool in
the designing of the wastewater treatment plants for the optimisation of their behaviour.
The evaluation of the pressure and velocity of the flow in these basins provides very
useful information about the flow properties. The data about the streamlines and
velocity field distribution allows us to know where the main recirculation regions are
taking place. This information will be priceless for the purpose of obtaining the
geometrical parameters of the basins in order to achieve a better performance for the
treatment plant. The obtaining of this optimum geometry will allow for a further
recirculation, if the energy losses are required; or will enable its avoidance if unwanted,
modifying in this way the detention times within the basin. The velocity and pressure
fields also provide invaluable information about the distribution of the discharge among
the outlets, which again can be redefined in order to improve the behaviour of the plant.
Thanks to the information obtained by this numerical evaluation of the flow, the water
treatment basins and channels can consequently be designed to fit the requirements of
the processes being carried out.
REFERENCES
Armaly B.F., Durst F., Pereira J.C.F. and Schönung B., Experimental and
theoretical investigation of backward-facing step flow. J. Fluid Mech. 127 (1983),
pp 473-496.
25
Babuska I., Error bounds for finite element method, Numerische mathematic 16
(1971), pp 322-333.
Baker A.J., A finite element computational theory for the mechanics and
thermodynamics of a viscous compressible multi-species fluid. Bell Aerospace
Research Rept. 9500-920200 (1971).
Benim A. C. and Zinser W., A segregated formulation of Navier-Stokes equations
with finite elements. Comput. Methods Appl. Mech. Engrg. 57 (1986) 223-237.
Brezzi F., On the existence, uniqueness and approximation of saddle point problems
arising from Lagrange multipliers. Rev. Française Automatique Informatique
Reserche Operationnelle, Ser. Rouge Anal. Numér., 8 (R2) (1974), pp 129-151.
Brooks A.N. and Hughes J.R., Streamline Upwind / Petrov-Galerkin formulations
for convection dominated flows with particular emphasis on the incompressible
Navier-Stokes equations. Comput. Methods Appl. Mech. Engrg 32 (1982), pp 199259.
Carey G. and Oden J., Finite Elements. Prentice-Hall (1984).
Chadwick A. and Morfett J., Hydraulics in Civil Engineering. Allen & Unwin
(1986).
Choi H.G. and Yoo J.Y., Streamline upwind scheme for the segregated formulation
of the Navier-Stokes equation. Numerical Heat Transfer 25-Part B (1994), pp 145161.
Choi H. G., Choi H. and Yoo J.Y., A fractional four-step finite element
formulation of the unsteady incompressible Navier-Stokes equations using SUPG
and linear equal-order element methods. Comput. Methods Appl. Mech. Engrg. 143
(1997), pp 333-348.
Cruchaga M. A. and Oñate E., A finite element formulation for incompressible
flow problems using a generalized streamline operator. Comput. Methods Appl.
Mech. Engrg. 143 (1997), pp 49-67.
Donea J., Giuliani S., and Laval H., Finite Element solution of the unsteady
Navier-Stokes equation by a fractional step method. Comput. Methods. Appl. Mech.
Engrg. 30 (1982), pp 53-73.
Espert V., García M., Sancho H. and López A., Modelo matemático
bidimensional para el estudio de flujo de agua a través de un decantador rectangular
con lamelas. Ingeniería del Agua 3 (1996) pp 16.
Ferziger J.H. and Peric M., Computational methods for fluid dynamics. Springer
(1996).
Fortin M., Analysis of the convergence of mixed finite element method. RAIRO
Anal. Numer. 11 (1977), pp 341-354.
Franca L. P. and Frey S. L., Stabilized finite element methods: II The
incompressible Navier-Stokes equations. Comput. Methods Appl. Mech. Engrg 99
(1992), pp 209-233.
Gartling D.K., Finite element analysis of viscous incompressible flow. Ph.D.
dissertation. University of Texas at Austin, Austin (1974).
Ghia U., Ghia K. N. and Shin C.T., High Re solutions for incompressible flow
using the Navier-Stokes equation and the multigrid method. J. Comput. Phys. 48
(1982), pp 387-411.
Hannani S.K., Stanislas M. and Dupont P., Incompressible Navier-Stokes
computations with SUPG and GLS formulations- A comparison study. Comput.
Methods Appl. Mech. Engrg 124 (1995), pp 153-170.
26
Haroutunian V., Engelman M. S. and Hasbani I., Segregated finite element
algorithms for the numerical solution of large scale incompressible flow problems.
Int. J. Numer. Meth. Fluids 17 (1993), pp 323-348.
Heinrich J.C., Huyakorn P.S., Mitchell A.R. and Zienkiewicz O.C., An upwind
finite element scheme for the two dimensional convective transport equation. Int. J.
Numer. Methods Engrg 11 (1977), pp 131-143.
Heinrich J.C., and Marshall R.S., Viscous incompressible flow by a penalty
function Finite Element Method. Comput. Fluids 9 (1981), pp 73-83.
Hughes T.J.R., Liu W.K. and Brooks A., Review of Finite Element analysis of
incompressible viscous flow by the penalty function formulation. J. Comput. Phys.
30 (1979), pp 1-60.
Jamet P., and Raviart P.A., Numerical solution of the stationary Navier-Stokes
equation by Finite Element Methods. Computing methods in applied sciences and
engineering (Ed. By Glowinski and J.L. Lions). Springer, Berlin (1973), pp 193223.
Kim S.W., A fine grid Finite Element computation of two-dimensional high
Reynolds numbers flows. Computers and Fluids 16-4 (1988), pp 429-444.
Kondo N., Third-Order upwind finite element formulations for incompressible
viscous flow problems. Comput. Methods Appl. Mech. Engrg. 93 (1991), pp 169187.
Kwak D., and Chang J.L.C., A three dimensional incompressible Navier-Stokes
flow solver. Part I. –INS3D Code. CFD Workshop. University of Tennessee Space
Institute, Tullahoma, TN (1985).
Ladyzhenskaya O., The mathematical theory of viscous incompressible flow.
Gordon & Breach (1969).
Laval H. and Quartapelle L., A fractional-step Taylor-Galerkin method for
unsteady incompressible flows. Int. J. Numer. Meth. Fluids 11 (1990), pp 501-513.
Metcalf & Eddy INC., Ingeniería de aguas residuales, tratamiento, vertido y
reutilización. Mc Graw Hill (1995).
Oden, J.T., Finite Elements of non-linear continua. McGraw Hill, New York
(1972).
Patankar S.V., Numerical heat transfer and fluid flow. McGraw-Hill (1980).
Ramaswamy B., and Jue T.C., Some recent trends and developments in Finite
Element analysis of incompressibe thermal flow. Int. J. Numer. Methods Engrg. 35
(1992), pp 671-707.
Rice J.G. and Schnipke R. J., An equal-order velocity pressure formulation that
does not exhibit spurious velocity modes. Comput. Methods Appl. Mech. Engrg. 58
(1986), pp 135-149.
Rodi W., Turbulence models and their application in hydraulics: A state of the art
rewiew. Balkema, Rotterdam (1993).
Sampaio P. A. B. de, A Petrov-Galerkin formulation for the incompressible NavierStokes equations using equal order interpolation for velocity and pressure. Int. J.
Numer. Meth. Engrg 31 (1991), pp 1135-1149.
Schneider G.E. Raithby G.D., and M.M. Yovanovich, Finite Element solution
procedures for solving the incompressible Navier-Stokes equations using equal
order variable interpolation., Numer. Heat Trans. I (1978), pp 433-451.
Shaw C.T., Using a segregated finite element scheme to solve the incompressible
Navier-Stokes equations. Int. J. Numer. Meth. Fluids 12 (1991), pp 81-92.
27
Shen S. F.and Habashi W., Local linearization of the Finite Element methods and
its application to compressible flow. Int. J. Numer. Methods Engrg. 10 (1976), pp
565.
Sohn J.L., and Heinrich J.C. A Poisson equation formulation for pressure
calculations in penalty Finite Element models for viscous incompressible flows. Int.
J. Numer. Meth. Eng. 30 (1990), pp 349-361.
Taylor C. and Hood P., A numerical solution of the Navier-Stokes equation using
FEM technique . Compt. &Fluids I (1973), pp 73-100.
Temam R., Navier-Stokes equations. North Holland, Amsterdam (1977).
Toit C.G. du, Finite element solution of the Navier-Stokes equations for
incompressible flow using a segregated algorithm. Comput. Methods Appl. Mech.
Engrg. 151 (1998), pp 131-141.
Turner, M., Clough R., Martin H., and Topp L., Stiffness and deflection analysis
of complex structures, J. Aero Sci. 23/ 9 (1956), pp.805-823.
Vellando, P.(Pablo Rodríguez-Vellando Fernández-Carvajal). On the resolution
of the Navier-Stokes equations by the Finite Element Method using a SUPG
stabilization technique. Application to some wastewater treatment problems. Dotoral
thesis. Universidad de La Coruña (Spain), Marzo, 2001.
Versteeg H.K. and Malalasekera W., An introduction to computational fluid
Dynamics. Harlow, UK (1995).
Weiyan T., Shallow water hydrodinamics. Elsevier (1992).
Zienkiewicz O.C., Constrained variational principles and penalty function methods
in Finite Element analysis, in Lecture Notes in Mathematics, Conference of the
Numerical Solution of Differential Equations, ed. G. A. Watson. Springer-Verlag,
Berlin (1974), pp 207-214.
Zienkiewicz O.C., R.H. Gallagher and P. Hood, Newtonian and non-Newtonian
viscous incompressible flow. Temperature induced flows. Finite element solution,
J.R. Whiteman, ed. The Mathematics of Finite Elements and Applications II
(MAFELAP 1975). Academic Press, London (1976).
NOTATION
Aυ
K ip
m
M
Mυ
n
Viscous coefficient matrix
B
Gradient of pressure matrix
Penalty matrix
Bε
bi
Prescribed velocity
Cυ (u, v ) Convective coefficient matrix
eξi , eηi Unit vectors in the ξ and η directions
f
fi
g
h
h
h*
h,*j
hξ , hη
k ij
k ij
N
Ni
p
p
p
q
Q1P0
r
Re
Rh
S0
Sf
t
Body force vector
Body forces components
Gravity force
Depth vector
Depth or size of the grid
Star depth
Characteristic lengths along ξ and η
Star gradient of depths
Diffusion
Artificial diffusion
28
Pressure-velocity coupling coefficient
Mass
Number of pressure nodes
Unsteady coefficient matrix
Manning roughness coef. or outward
unit vector normal to the interface
Number of velocity nodes
Velocity shape function
Pressure vector
Pressure
SUPG weighting term
Weighting function
Bilinear velocity-constant pressure BE
Inertial relaxation factor
Reynolds number
Hydraulic radius
Geometric slope
Friction slope
Time
ti
u
u
u~i
v
v
v~i
Velocity along the y direction
Velocity vector along the y direction
Pseudo-velocity in the y direction
αp
Weighting functions
Pressure relaxation parameter
wi
η
µ
Traction vector
Velocity along the x direction
Velocity vector in the x direction
Pseudo-velocities in the x direction
ν
ξ
ρ
σ ij
τ ij
υ
χi
Velocity relaxation parameter
αu
α ξ ,αη Reynolds numbers along ξ and η
Γ
ε
Ω
∂Ω
Boundary of the Ω domain
Penalty parameter
29
Local spatial variable
Dynamic viscosity of the fluid
Kinematic viscosity of the fluid
Local spatial variable
Density of the fluid
Stress along the boundary
Stress
Velocity vector
Pressure shape function
Domain of integration
Domain boundary