INTERNATIONAL JOURNAL FOR NUMERICAL METHODS I N FLUIDS, VOL.
6,241-253 (1986)
zyxwvu
zyxwv
zyxwvutsrqp
MODELLING A RECIRCULATING DENSITY-DRIVEN
TURBULENT FLOW
BRUCE A. DEVANTIER
Department of Civil Engineering and Mechanics, Southern Illinois University, Carbondale, I L 62901, U.S.A.
AND
BRUCE E. LAROCK
Department of Civil Engineering, University of California, Davis, C A 9561 6, U.S.A
SUMMARY
zyxwvu
A mathematical model of turbulent density-driven flows is presented and is solved numerically. A form of the
k--E turbulence model is used to characterize the turbulent transport, and both this non-linear model and a
sediment transport equation are coupled with the mean-flow fluid motion equations. A partitioned, NewtonRaphson-based solution scheme is used to effect a solution. The model is applied to the study offlow through a
circular secondary sedimentation basin.
KEY WORDS
Finite-element model Turbulence Density flow
INTRODUCTION
Fluid motions which are initiated or significantly influenced by gradients in fluid density are called
density currents. On a large scale both the earth’s gravitational field and its rotation influence fluid
motions when less dense fluid is at a lower potential level than more dense fluid. Fortunately the
rotational effect becomes insignificant on smaller scales. Fluid density differences may be caused
either by differences in temperature or by the presence of contaminants, such as salt or suspended
solids. The prediction of flow patterns in such density-driven flows must often simultaneously cope
with the facts that the flow is predominantly turbulent and that the structure of the turbulence is
itself influenced by the non-constant density field. In confined regions these effects may lead to the
creation of zones where the flow turns back on itself or recirculates. The computation of the
behaviour of a recirculating density-driven flow is usually complicated further because the flow is
driven by spatially continuous density gradients and is not amenable to modelling with the
assumption of the presence of distinct density layers, as can happen when the flow field is infinite.
Finally, such flows are dramatic examples of a situation where a small deviation from uniformity
(in density) causes not a small but a large behavioural change in related phenomena (the flow field),
contrary to one’s usual expectation.
This paper considers the modelling of what is essentially a turbulent density current driven by
the presence of a distribution of suspended solids of higher density than the carrier fluid (water); the
settling of the solids creates spatial gradients in the bulk fluid density (solids plus fluid). A higherorder turbulence closure model, the k--E model, is used to model the turbulent transport effects. The
use of simpler models would result in a much simpler mathematical form for the final equation set,
zyx
zyx
zyxwvu
zyxwvuts
0271-2091/86/040241-13$06.50
0 1986 by John Wiley & Sons, Ltd.
Received 8 February 1985
Revised 10 October 1985
242
zyxwvutsrq
zyxwvut
9. A. DEVANTIER A N D 9. E. LAROCK
but those models would require a priori knowledge of the nature of the turbulent transport. For
strong density currents the turbulent transport is intimately linked to a flow pattern which is
determined primarily by the density distribution. Consequently a more general turbulent flow
closure model is preferred even though it is computationally more complex.
The finite element method is used to solve approximately the resulting set of coupled partial
differential equations, owing to the ease with which boundary conditions can be implemented, and
also to the ease with which the computational grid spacing can be modified. The model equations
are therefore presented directly in integral form. No numerical upwinding is employed in the
solution of the equations, so the proper choice of grid spacing is important.
A practical application of the model to the study of flow through a circular secondary clarifier,
which is a common waste-water treatment process, will also be presented. In this way the
capabilities of the model, its operational features and (good and bad) characteristics can be
displayed.
MODEL EQUATIONS
zyxw
zyx
A set of equations that suitably model a sediment laden, density-affected turbulent flow has
recently been developed and presented by DeVantier and Larock.' The model is based on laws of
mass, volume and momentum conservation for the transporting fluid (usually water) and the
transported, suspended solid particles. The use of local volume averaging and the usual Reynolds
decomposition of variables in turbulent flow produces a continuum model that is appropriate for
application of multidimensional flows. For ready reference the partial differential equations which
make up the model are listed in Appendix I.
For steady two-dimensional radial flow the element-level residuals that result from application
of the Galerkin finite element method to the mean flow transport equations are
F,
1.
Mi( &(rU)
=
jQeNi[r( U g + W aZ
~ + a ar
p)_w]dR-
F,=
F,
+
=
IQe
aw + W-aw + ap + qAAVEg
aZ aZ
zyxwvu
zyxwvut
Nir( Uar
F,=JQeNir(Ug+
+
j Q e r ( E ~ + aZ
W ~ ) d R
1
Q.
-
W ~ ) d R + ~ Q ~ r A ( l - aNi
h)V,dn
aZ
aZ
rE, (aAaNi
-~
+--aAaNi) dR
ar ar
aZ aZ
Following consistency arguments advanced by King2 the density of the bulk fluid (the weighted
zyxwvut
zyxwvutsr
zyxwvutsrqponmlkjihg
zyxwvu
zyx
zyxwvut
243
RECIRCULATING DENSITY-DRIVEN TURBULENT F L O W
density of the fluid and the transported sediment) must have a single value for each element. Hence
a single average value of A, A A v E , appears in the body force term v ] l \ A v & of the vertical momentum
equation (2). The factor v] = (p, - pf)/pf is the normalized density difference between the solid of
density p, and fluid of density pf, so the body force term represents only the body force associated
with the sediment fraction. The modified pressure P at a point is the difference between the actual
pressure and the clearwater hydrostatic pressure at that point. The two co-ordinates are the
horizontal radial direction rand the vertical direction z in cylindrical co-ordinates. F,, F,, F , and
F A are the element level residuals for the modified pressure P, radial velocity U , vertical velocity W ,
and sediment volume fraction A, respectively. Integration is performed over an element domain Re
having an element boundary re,and I, and I, are the components of the unit outer normal from the
boundary. A mixed weighting of the residuals has been chosen by analogy to the usual treatment of
the Navier-Stokes equations. Hence N i is the eight-node biquadratic function that is used to
interpolate all variables except P and weight the momentum and mass transport equations; M i is
the bilinear function that is used to interpolate P and weight the continuity equation.
As equation (4)shows, the transport of the sediment is influenced not only by convection by
the fluid mean velocity components U and W but also by two parameters characteristic of the
sediment motion itself; they are the sediment settling velocity V, and the turbulent diffusion
coefficient E,.
The other notable transport terms in equations ( 2 ) and ( 3 ) are the turbulent momentum flux or
velocity correlation or ‘stress’ terms, the overbarred terms; they are also called Reynolds stresses
(per unit mass) and must be suitably represented in terms of other variables to close the equation
set. The viscous contributions to the fluid stress are assumed to be negligible in this turbulent
flow. The terms in these equations are all non-dimensional, based on a characteristic velocity
U,, height h and the clearwater kinematic viscosity v and density pf.
Many turbulence models have been proposed; they range from the simple to the very complex.
The simpler models assume local equilibrium between turbulence generation and dissipation and
neglect history and transport effects. The more advanced models also consider the latter effects
which are believed to be significant in the present study. The turbulent stress terms have therefore
been modelled via the k--E turbulence model proposed by Rodi3 for density-affected flows. The
model is built on the kinematic eddy viscosity constitutive relation, which in cylindrical
co-ordinates may be written
zyxw
zyxwvu
-
uu = $k
au
2~,--,
ar
U
uv = #k - 2 ~ , - - ,
r
.-
+
-
aw
-
ww = 3k - 2 ~ , - - ,
aZ
IC\
IJI
-
Here k = (G+ uu ww)/2 is the turbulence kinetic energy per unit mass. The kinematic
eddy viscosity v, is expressable by dimensional analysis in terms of k and its rate of dissipation
E as
V , = C,k2/&,
(6)
where c, is a model constant.
Final closure of the equation set is accomplished by developing transport equations for k and
8. These equations are listed in Appendix I. The Galerkin residuals of the resulting equations are
ak
ak
U-+W--Pr+E-B
ar
aZ
)
dQ+
,,(akaNi akaNi)
r- -----+-dQ
J- ok ar ar
aZ aZ
244
and
zyxwvutsrq
zyxwvu
zyxwvutsrqpon
zyxwvuts
zyxwvuts
zyxwvu
B. A. DEVANTIER A N D B. E. L A R O C K
aE
a&
E
U p + W--cc,,-[Pr+(l
ar
aZ
k
-cE3)B]
The model constants cZ1= 1.44, cE2= 1.92, ce3= 0.8, cc,= 1.3, ok= 1.0, and c, = 0.09 are
given by RodL3 In addition, E , = yv, with y = 1.2 is chosen.’ In these equations Pr and B are
the production of turbulence kinetic energy by the mean shear rate and by buoyancy and take
the forms
Pr = v , [ 2(
g)2+ (g+
+ 2(
2) :
g) +
2(
r>]
and
Several of the boundary conditions that are required to complete the model require some
explanation. For boundaries where solid walls confine the flow domain, the velocity can be
determined once the wall shear stress zw is known. The wall shear is related to the near-wall
velocity U , by a logarithmic law for wall-bounded turbulent flows
where u* = ( ~ ~ / p is) ”the~ shear velocity, ti is the von Karman constant, and v is the kinematic
viscosity. The wall shear stress, which is determined by solving equation (11) for , ,z is then
used for the boundary integral forms of shear stress in equations (2) and (3). The boundary of
the computational region is displaced a small distance a from the wall to ensure that flow in
the computational region is indeed turbulent so that the k--E model is valid. The constant K is
9.0 for hydrodynamically smooth walls. The standard k--E model then assigns wall values to the
turbulence parameters as
and
zyxwvuts
zyxwv
El,
4
= -.
tia
Fluid free surface boundaries are treated as planes of symmetry for velocities and the turbulence
variables k and E. Along a free surface there is no sediment flux, so
+
+
yv,( gl, El,) A(1 - A)Vs12= 0.
A similar relation is applied along wall boundaries, where a fraction
the wall is resuspended by the local turbulence; thus
yV,
(g +
1,
1,)
+ pA (1
-
A) V,1,
= 0.
p of the sediment reaching
RECIRCULATING DENSITY-DRIVEN TURBULENT FLOW
EQUATION SOLUTION TECHNIQUES
zyxw
zy
245
It is not difficult to see that the model governing equations are strongly coupled and non-linear.
A Newton-Raphson scheme for solution of the equations is chosen because of its potentially
quadratic convergence rate. The method is applied in the form
zyxwvuts
zyxwvu
zyxwvuts
zy
zyxwvut
where F iis a residual and q j is an array of nodal values of the unknowns. These residuals F i
are global nodal residuals that are assembled by adding together all of the individual element
contributions in the usual finite element way. The full incremental change Aqj for the nodal
variables is not necessarily imposed; instead the unknowns are updated via
qyew= qoy + aAqj.
(17)
The choice for a in equation (17) should be made so that the nodal unknowns converge to a
solution in some optimal fashion. Following Fletcher: a line search was implemented to ensure
a decrease in an objective function which is representative of overall solution convergence. The
extreme non-linearity of the equations did not allow the use of more powerful search techniques,
so a simple interval halving scheme was employed in conjunction with an objective function
defined as the sum of the absolute values of the nodal residuals. A further limitation upon CL
was imposed because the variables k, E and A must by definition be non-negative. The limitation
that precludes negative values of certain variables is known as a realizability constraint. Thus
the realizability constraint was first satisfied before a was determined by optimization, and the
upper limit for a during the optimization was defined by the realizability limit on a.
The earliest attempts to obtain solutions of the equations applied the Newton-Raphson
scheme directly to the full equation set, but these attempts were not successful. Consequently a
partitioned matrix approach similar to that proposed by Schamber and Larock' was developed
and used. The method sacrifices the convergence rate of the full Newton-Raphson method in
return for a more stable convergence by partitioning the equations into two sets and applying
the Newton-Raphson scheme separately to each while holding the other set constant. This
approach was deemed necessary because the values of the turbulence parameters k and E could
not reasonably be estimated without prior knowledge of the flow field. The structure of the flow
field is in turn strongly influenced by density differences that are affected by the sediment transport
equation. The partitioned matrix approach turned out to be a method that is less sensitive to
relatively poor initial estimates for k and E.
The best choice for the partitioning of variables was not immediately obvious. Partitioning
the variables into the subsets ( U j ,W j ,P j ) and ( k j ,c j , A j ) leads to subsets of equations of roughly
equal size and was initially attempted. This grouping was not acceptable because of the strong
dependence of P upon A. By operating on these variables in separate subsets, the pressure P
(above clearwater hydrostatic pressure) could not react to changes in excess body force caused
by new estimates for A, and the solution scheme diverged. The successful partitioning of the
variables is therefore ( U j ,W j ,P j , A j ) and ( k j ,cj). These two groups then alternately become the
qj array in equations (16) and (17), with one set held constant while the other is operated upon
by the Newton-Raphson scheme.
The first set of iterations is performed upon the mean flow variables ( U j ,W j ,P i , Aj)while holding
the turbulence variables ( k j ,E ~ )constant. This step requires the prescription of an initial
distribution of v, without a knowledge of k, and so the 2k/3 contribution to the normal stresses
iiii and WW were set to zero. Subsequent sets of iterations on the mean flow variables included
the 2k/3 contributions, once the turbulence variables had been operated upon for a set of iterations.
zyxwvut
zy
246
zyxwvutsrq
zyxwvut
zyxwvut
zyx
zyxwv
zyxwvuts
zyx
B. A. DEVANTIER A N D B. E. LAROCK
The first sets of iterations on the ( k j ,c j ) subset were computed with the use of modified forms
of the Newton coefficient matrix dFi/aqj to avoid divergent results caused by the strong
non-linearities in the two governing equations. There are two major ways in which the
non-linearities affect the convergence of the Newton estimates:
1. The first cause is the behaviour of the sink term involving E2/k in the E transport equation.
This functional dependence causes the Newton estimates to diverge rapidly from the solution
whenever the E estimates are incorrect and especially when the E estimates are poor. To avoid
this behaviour, the Newton coefficient matrix contributions that are associated with the E
equation are formed while assuming for the sink term that &/kis a locally specified constant.
2. The second cause is the k2/&dependence of v,. The turbulent viscosity v, appears throughout
both the k and E equations in both the terms containing Pr and the diffusive terms. When v, is
assumed to be a locally specified constant in forming the Newton coefficient matrix, the
iterates converge more smoothly although more slowly.
With these assumptions the Newton matrix contributions for F , are
dF,MODI
akj
=-
1.
&
rNiNjc,,&Pr
+(1
-
c E 3 ) BdQ,
]
(18)
The boundary integrals that appear in F , in equation (8) do not appear either because the
boundary conditions cause the integrals to be zero owing to flux considerations or because k
and E are specified on the boundary, and the entire nodal residual equation is omitted from the
set. Note that no simplification is made for the source term c,,(&/k)Pr because the factor v, in
Pr causes the overall term to depend linearly on k. The superscript MOD1 denotes that equations
(18) and (19) are the first modified forms of the coefficients. They are used until the E estimates
are sufficiently well behaved that the full non-linear nature of the c,,c2/k sink term can be
tolerated. The second modified form, MODII, is then employed in the forms
and
aF
,MOD"
- aF,MOD'
aEj
aEj
+
&
r N i NjcEzidQ.
Finally, the full Newton form which considers the non-linearity of vt is used; for example,
a F y L - aF,MoD"
aEj
aEj
The use of these modified equation forms will be described further in the next section.
MODEL RESULTS
Radial flow circular sedimentation basins are used extensively throughout the world for both
primary and secondary sewage treatment. In some primary treatment units the solids or sediment
247
RECIRCULATING DENSITY-DRIVEN TURBULENT FLOW
zyxwvuts
zyxwvutsr
zyx
zyxwvut
concentrations are sufficiently low to allow the sediment transport to be considered independently
of the determination of the flow field.5In most practical applications this is not the case, however,
and this is particularly true of secondary clarifiers. Indeed the current mathematical model was
developed so that it would aid in the understanding of secondary clarifiers.
Figure 1 is a radial-section schematic diagram of a centre feed radial flow secondary clarifier.
The sediment-bearing fluid enters the centre of the basin, is clarified by gravity sedimentation
as it flows outward, and most of the fluid exits over the eMuent weir. The sediment, which is
deposited on the basin floor, is periodically removed by a slowly rotating suction apparatus (not
shown) which also removes some of the fluid.
A circular basin was chosen for a number of reasons. (1) These basins are by far the most
common design configuration used in the United States. (2) The flow ought to be more nearly
two-dimensional than the flow in rectangular basins; in rectangular basins both corners and
side walls encourage the development of secondary currents, and these geometric effects are absent
here. (3) The velocities at the point sink model of the eMuent weir are not nearly so large as
those at the exit from a rectangular basin, owing to radial deceleration, and thus they should
affect the flow field only locally. It is regrettable that no suitable data exist on solids-induced
density currents in circular basins,6 but qualitative comparisons can still be made. Currently
there is considerable interest in the measurement of the details of flows in settling basins, but
practical problems such as probe fouling, the avoidance of sludge removal mechanisms, and the
very low velocities themselves make the collection of such data very difficult. The authors are
aware of one measurement programme in progress at the University of California, Davis, which
seeks to collect mean flow and turbulence data. It is hoped that this paper might encourage
additional investigators to obtain detailed flow measurements in operating circular basins.
This model was used to determine flow patterns and sediment concentration patterns in such
clarifiers. The final computational mesh for the problem is shown in Figure 2. The line r = 0 is
z
t
PHYSICAL DOMAIN
Figure 1. Schematic diagram of radial flow clarifier cross-section
zyxw
1.0
0.0
1.o
2.0
3.0
4.0
Figure 2. Computational mesh. Vertical and horizontal lengths in basin depths (125 ft, 3.8 m)
248
zyxwvutsrq
zyxwvutsrq
zyxwvutsrqp
zyxwvu
zyx
zyxwvu
zyx
B. A. DEVANTIER A N D B. E. LAROCK
the axis of symmetry for the basin, and the computational region begins 0.65 basin depths away
from r = 0 just outside the baffle. The baffle creates a nearly uniform inlet profile at the edge of
the computational region. The lines z = 0.0 and r = 4.0 coincide with wall boundaries, and thus
computation begins a small distance inward from the walls. Outflow over the effluent weir at
the upper right corner ( r = 4.0, z = 1.0) is locally modelled as potential flow towards a sink in
the corner. The strength of the sink then defines the outflow velocities near that corner. In the
presumed absence of wind across the basin it is assumed that the vertical cross-section is
representative of the basin, and thus that the flow may be modelled as two-dimensional.
The grid shown in Figure 2 was the result of a number of grid refinements. The most important
of these refinements was above the inlet region near z = 1.0, a free surface. The solution shows
that this is a region where a return current takes a sharp downturn, and k and E values change
drastically radially here. Owing to an insufficiently fine mesh, the k k iterates locally diverged
at a few nodes near the free surface. The divergence remained localized because one or two free
surface nodal values of k were being forced toward negative values so strongly that the realizability
constraint allowed only very small relaxation factors, and thus the iteration algorithm ground
to a halt. The refined grid of Figure 2 contains 260 elements and 853 nodes with 2662 U - W - P - A
equations and 1530 k--E equations.
The proper initialization of variables (as in any Newton scheme) was very important for this
problem. The values of k and E were initialized as decreasing linearly in the r direction. This
produces a linearly decreasing v, which is what might be expected as a consequence of the linear
deceleration caused by the radially increasing cross-section. Initialization of the velocity
components to zero caused no algorithmic difficulties, but the values of A and P must
be compatible to avoid immediate divergence of the algorithm. This agreement is required
because the gradients in P which drive the flow are three orders of magnitude smaller than most
of the values of P itself. Thus the deviation in P from the clearwater hydrostatic pressure must
primarily balance the excess density body force VAg but also secondarily provide the driving
force for the mean flow via its gradient. With this in mind, a linearly decreasing profile in z was
chosen in initializing A; then the equation
was integrated to obtain P, and the constant of integration was chosen to agree with the chosen
reference value for P.
A vector representation of a velocity field is shown in Figures 3 and 4. The inflow velocity to
the computational region is 0.396cm/s, and the inlet concentration of solids is 1400mg/l. The
low velocity is typical of sedimentation basins. However, the flow is turbulent in spite of the
low velocity because the length scale is so large (basin depth is 3.8m). The fluid withdrawal
velocity along the bottom of the basin is not noticeable because it is 4.6 per cent of the inflow
velocity, and yet this results in the removal of 33 per cent of the inflow from the basin. The
density current created by the sediment concentration gradients is clearly evident in the Figure.
There is a strong current along the bottom and a return current near the top of the basin. This
flow pattern is similar to those measured by Larsen’ in rectangular basins. The return current
cannot be discerned in the outer portion of Figure 3 because the velocities are very low in
comparison with the bottom current and thus appear as x s. Figure 4 presents a magnified view
of the velocity field in the outer region. A stagnation point that divides the return flow and the
exit flow is also apparent near the upper right corner.
The sediment concentration distribution which primarily drives the flow is shown in Figure 5.
Clearly the inlet region is subject to relatively strong density gradients. The 500mg/l contour
Velocity scale
zyxw
zyxwvutsr
zyx
RECIRCULATING DENSITY-DRIVEN TURBULENT FLOW
249
zyxwv
C,, = 1400mg/L
Lwy
0.0 Olft/s (003 m/s)
-~
~
-
*
+ / + +
* / +
+
/
~
zyxwvutsrqponmlkjihgfedcbaZYXWVUTSRQPONMLKJIHGFEDCBA
:
:
,“.
x
L
A
4
+
4
4
t
+
+
+
4
A
X
X
x
x
x
x
X
x
x
x
x
4
x
x
x
x
4
x
x
x
x
x
x
x
x
zyxw
zyxwvutsrq
f
X
x
X
*
+
+
*
x
+
+
x
*
+
+ +
x
x
*
*
* *
Figure 3. Plot of predicted velocity vector field (small velocity scale). C , , = 1400mg/l
Velocity scale
,
I
0.0
I
I
I
I
0.05 ft/s (0.016m/s)
zyxwvu
zyxwv
Figure 4. Plot of predicted velocity vector field (expanded velocity scale).
130
200
Contours in mg/L
Figure 5. Contour lines of predicted suspended solids concentration in mg/l
islands near the end wall are an anomaly that is the result of an instability in the A-field
prediction near the corner stagnation point. Steep concentration gradients near the corner cannot
be properly resolved by this particular grid. Further mesh refinement was not attempted owing
to computer core storage limitations and also because the instability disappeared in subsequent
solutions for flow with lower inflow sediment concentrations and accordingly lower underflow
current velocities. The behaviour is localized, and the overall effect on concentration prediction
is minimal.
250
zyxwvutsrq
zyxwvu
zyxwvutsr
zyxw
B. A. DEVANTIER A N D B. E. LAROCK
Solutions for the k- and &-fieldswere only obtained after an important modification was made
in the transport equations given in equations (7) and (8). When the buoyancy term B was included
in these equations, no solutions could be obtained for k and E for an initially specified velocity
and sediment concentration field. This occurred because the basin was stably stratified and B
became a net sink of k. Its value was so large that it swamped the shear production Pr and
drove values of k and E toward negative values. Turner' notes that it is a physical impossibility
for ( P r + B ) to act as a net sink. Turner also cites a number of sources which assert that
-B/Pr, the flux Richardson number, has some theoretical limit less than unity, at which
turbulence collapses. He concludes from these sources that the maximum Richardson number
should lie in the range 0.1 to 0.3. Therefore buoyancy damping should have only a secondary
effect upon values of k and E, and thus upon v,. Larsen,' as well as others, has consistently
found strong turbulence in secondary sedimentation basins in normal operation. Consequently,
the present numerical solutions have been obtained with B arbitrarily set to zero, with the
thought that such a solution should still capture the basic features of the turbulence.
Although the initial solution to the problem was difficult to obtain, the development of
solutions to subsequent related problems was substantially easier and less time consuming. In
those cases the solution to the first problem could be used as the first estimate to start the
iterative solution procedure for a subsequent problem. Solutions for problems involving lower
inlet solids concentrations were thus bootstrapped successively from one to the next. Each
subsequent solution could be viewed as a solution obtained by continuation in A-space or as
an incremental unloading. The inflow concentration of sediment was the primary factor
determining the strength of the bottom current, so that each lower inlet concentration solution
converged slightly faster than the previous one. A comparison of computation schedules for the
initial solution and a subsequent solution (see Appendix 11) reveals that the bootstrap solution
required less than one third the number of iterative cycles (14 iterations vs. 52). Both solutions
required more than twice as many k - E iterations as U - W - P - A iterations, because the k--E
equation set contained the strongest non-linearities. In general, a k--E iteration consumed
approximately one third the CPU time of a U - W - P - A iteration owing to the smaller number
of equations and simpler boundary conditions.
The iteration schedules shown in Appendix IT are only in part determined by the computational
algorithm. Within each set of iterations (i.e. U - W-P-A, k--E, MODI, etc.) the line search routine
chose c1, but the choice of the number of iterations within a set and the timing of the switches
from one k--E form to another were controlled by the operator. The results of each set were
examined before subsequent sets of iterations were performed so that convergence rates could
be monitored. The logic behind the choice of the iteration schedule was necessarily often intuitive,
and several subsets of iterations had to be rerun with different k--E modes or increased iterations
to avoid divergence. At this point it would not be a simple task, indeed it may not be possible,
to automate the iteration schedule completely. Further study is definitely needed in this area.
zyxw
zyxwvuts
zyxwvutsr
SUMMARY AND CONCLUSION
A two-dimensional model that captures the dominant features of a density-driven turbulent flow
has been presented. The Galerkin finite element method has been applied to this model; the
resulting non-linear equation set has been solved by partitioning the equations into two subsets
and employing the Newton-Raphson method (and two modified forms thereof). The solution
scheme has been aided by use of a line search technique. It has been further modified by the
need to satisfy a non-negativity constraint on k , E and A. The application of the model to flow
through a secondary sedimentation basin illustrates the features of the solution techniques.
RECIRCULATING DENSITY-DRIVEN TURBULENT FLOW
zyxw
25 1
From this work it is clear that work remains to be done in several areas. In the model itself the
role of the buoyancy production terms in the turbulence closure is not fully understood, and
additional research to resolve this behaviour is desirable. Further developments that contribute to
the reliability or robustness of the solution technique are desired; implementation of realizability
constraints and a line search method in a finite element solution algorithm are also features that
have received relatively little attention heretofore. It is felt that these techniques will become more
important as more research investigators tackle turbulent flow problems where the distribution of
a major driving force for the flow, here the creation of density gradients, is itself a part of the
solution (and a determinant of the turbulence properties) and is therefore unknown a priori.
zyx
ACKNOWLEDGEMENT
The major portion of this work was completed with the support of U.S. National Science
Foundation Grant CME-7914762, which the authors greatly appreciate.
zyxwvutsr
zyxwvu
zyxwvutsr
APPENDIX I
zyxwvut
The partial differential equations which describe the mean flow model and the turbulence model
are listed here in cylindrical co-ordinates. The individual variables are defined and explained in the
main text.
The mean flow transport equations are
a
-(rU)
ar
+ Y- aw
aZ = 0
(Z
r U-
(continuity),
+ W-Z +r + -(ruu)
) i r - -
-
uu
a+ r-(uw)
aZ
=0
(r momentum),
+ r]Ag = 0
( z momentum),
(sediment conservation). (27)
The k--E turbulence model transport equations are
+ r(-
Pr+E-B)=O
E2
E
+rc,,--rc,,-[Pr+(l
k
k
( k transport),
(28)
-c,JB]=O
(E
transport).
(29)
Equations (5),(6), (9) and (10) and the accompanying main text define these turbulence variables
further.
zy
252
zyxwvu
zyxw
zyxwvut
zyxwv
zyxwv
zyxwvu
B. A. DEVANTIER A N D B. E. LAROCK
APPENDIX 11. COMPUTATION SCHEDULES
Original solution, inflow concentration = 1400 mg/l
Iteration numbers
1, 2
3-6
7-13
14, 15
16, 17
18, 19
20, 21
22, 23
24, 25
26, 27
28-30
31, 32
33,34
35, 36
37, 38
39, 40
41,42
43
44-46
47
48,49
50
51, 52
Mode
U - W-P-A
v, = 009
U-W-P-A
V , = 0.09 -+ 0.00 1 1
k-& MODI
k-& MODII
k-& Full
U-W-P-A
k-& MODI
k-6 MOD11
k-& Full
U - W-P-A
k-& MODII
k-& Full
U - W-P-A
k-& MODII
k-& Full
U - W-P-A
k-& Full
U - W-P-A
k-& Full
U-W-P-A
k-& Full
U - W-P-A
k-& Full
Relaxation factors
0.6, 1.00
1.0, 1.0, 1.0, 1.0
0.9, 1.0, 0.9, 1.0, 0.9, 0.5, 0.8
1.0, 1.0
1.0, 1.0
1.0, 1.0
1.0, 1.0
0.9, 1.0
0.6, 1.0
1.0, 1.0
1.0, 1.0, 1.0
05, 1.0
1.0, 1.0
1.0, 1.0
1.0, 1.0
1.0, 1.0
1.0, 1.0
zyxwvutsrq
zyxwvutsrq
1.o
1.0, 1.0, 1.0
1.o
1.0, 1.0
1.o
1.0, 1.0
Bootstrap solution, inflow concentration = 11 20 mgll
Iteration numbers
1
2, 3
4, 5
6
7, 8
9
10, 11
12
13
14
Mode
Relaxation factors
U - W-P-A
MOD11
k-& Full
U - W-P-A
k-8 Full
U - W-P-A
k-& Full
U-W-P-A
k-& Full
U - W-P-A
1 .o
1.0, 1.0
1.0, 1.0
1.o
1.0, 1.0
1.o
1.0, 1.0
1 .o
1.o
1.o
k-8
zyxw
zyxwvu
zyxwv
zyx
zyxwvu
zyxwvutsrqp
RECIRCULATING DENSITY-DRIVEN TURBULENT FLOW
253
REFERENCES
1. B. A. DeVantier and B. E. Larock, ‘Sediment transport in stratified turbulent flow’, J . Hyd. Engineering, A S C E , 109,
(HY12), 1622-1635 (1983).
2. 1. P. King, ‘A linite element model for three dimensional flow’, Report 9150 to U S . Army Corps of Engineers. Vicksburg,
Mississippi; RMA Inc., Lafayette, California, 1982.
3. W. Rodi, Turbulence Models and Their Application in Hydraulics, IAHR Publication, Delft, Netherlands, 1980.
4. R. Fletcher, Practical Methods of Optimization: Unconstrained Optimization, vol. I , Wiley, New York, 1980.
5. D. R. Schamber and B. E. Larock, ‘Numerical analysis of flow in sedimentation basins’, J . Hyd. Div., A S C E , 107, (HYS),
575-591 (1981).
6. A. I. Stamou and W. Rodi, ‘Review of experimental studies on sedimentation tanks’, SFB 210/E/2, University of
Karlsruhe, F.R. Germany, August 1984.
7. P. Larsen, ‘On the hydraulics of rectangular sedimentation basins: experimental and theoretical studies’, Report N o .
1001, Dept. Water Resources Engr., Lund Inst. Tech., University of Lund, Sweden, 1977.
8. J. S. Turner, Buoyancy effects in Fluids, Cambridge University Press, Cambridge, U.K., 1973, pp. 38-164.