You are on page 1of 28

Computers & Fluids 32 (2003) 403–430

www.elsevier.com/locate/compfluid

One- and two-equation turbulence models for the prediction


of complex cascade flows using unstructured grids
D.G. Koubogiannis, A.N. Athanasiadis, K.C. Giannakoglou *
Laboratory of Thermal Turbomachines, National Technical University of Athens, P.O. Box 64069,
15710 Athens, Greece
Received 13 October 1999; received in revised form 4 June 2001

Abstract
One- and two-equation, low-Reynolds eddy-viscosity turbulence models are employed in the context of a
primitive variable, finite volume, Navier–Stokes solver for unstructured grids. Through the study of the
complex flow in a controlled-diffusion compressor cascade at off-design conditions, the ability of the models
under consideration to predict the laminar separation bubble close to the leading edge and the boundary
layer development is investigated. In order to control the unphysical growth of turbulent kinetic energy
near the leading edge stagnation point, appropriate modifications to the conventional models are employed
and tested. All of them improve the leading edge flow patterns and significantly affect the size of the pre-
dicted laminar separation bubble. The use of an adequately refined mesh around the airfoil, that is formed
by triangles placed in a quasi-structured way, allows for the generation of grid elements of moderate aspect
ratios. This helps to readily overcome any relevant problems of accuracy; a second-order upwind scheme
without flux limiters or least squares approximations is successfully employed for the gradients. The test
case includes quasi-3D effects by considering the streamtube thickness variation in the governing equations.
Ó 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Unstructured grids; Low-Reynolds eddy-viscosity models; Compressor cascade flows; Laminar separation
bubble; Turbulent production

1. Introduction

Two-equation turbulence models or competitive one-equation models, are considered to be the


minimum level of closure in Navier–Stokes solution methods that are capable to cope with

*
Corresponding author. Tel.: +30-1-772-1636; fax: +30-1-772-3789.
E-mail address: kgianna@central.ntua.gr (K.C. Giannakoglou).

0045-7930/03/$ - see front matter Ó 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 4 5 - 7 9 3 0 ( 0 1 ) 0 0 0 8 6 - X
404 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Nomenclature

AVDR axial velocity density ratio (ðqVax Þin =ðqVax Þout )


c blade chord
E total energy per unit volume ðEeff ¼ E  ðc  5=3Þqk=ðc  1ÞÞ
G generation of turbulent kinetic energy
k turbulent kinetic energy
KN ðPÞ set of neighbouring nodes to node P
KT ðPÞ set of neighbouring triangles to node P
n, nþ dimensional and non-dimensional wall distances
p pressure (peff ¼ p þ 23 qk)
Pk turbulence production term
P, Q grid nodes
Pr,Prt laminar and turbulent Prandtl number ðPr ¼ 0:72; Prt ¼ 0:9Þ
Re Reynolds number
oui ou
Sij mean rate of strain ðSij ¼ 12 ðox þ oxji ÞÞ
j pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e
S dimensionless mean rate of strain e S ¼ ðk=eÞ 2Sij Sij
t time
Tt turbulence time scale
Tu turbulence intensity
T,G grid triangle and its barycenter
u1 ,u2 Cartesian components of the velocity vector V
x1 ,x2 Cartesian coordinates
Greek Letters
c ideal gas constant ðc ¼ 1:4Þ
dij Kronecker delta
e turbulent dissipation per unit mass
h temperature
j von-Karman constant ðj ¼ 0:41Þ
l viscosity coefficient ðleff ¼ l þ lt Þ
q fluid density
s viscous stress
v thermal conductivity coefficient ðv ¼ cl=RePr; veff ¼ v þ vt Þ
x specific dissipation rate
X vorticity magnitude X ¼ jrotVj
oui ou
Xij vorticity (Xij ¼ 12 ðox  oxji Þ)
j pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e
X dimensionless vorticity X e ¼ ðk=eÞ 2Xij Xij
Subscripts
i,j,k Cartesian directions ð¼1; 2Þ
ax axial
eff effective quantity
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 405

in inflow
out outflow
s tangent to the wall
t turbulent
w wall

complex flows. In the past, this was the outcome of many similar studies (for example [1,2], to
mention only those dealing with the same application). Two-equation eddy-viscosity models have
been credited certain advantages, mostly related to their simplicity and their superior numerical
properties, compared to other more sophisticated models, like those based on Reynolds stress
closures. In their conventional forms, two-equation models often yield poor predictions for flows
that encounter adverse pressure gradients, curvature, rotation or complex strain fields. In order to
overcome these deficiencies, without resorting to more sophisticated and more expensive turbu-
lence models, various modifications have been proposed in the literature. Among them, those
employed in the present method, are recapitulated in this text.
Working even with two-equation models, the low-Reynolds terms which are activated close to
solid walls are a source of convergence difficulties. For structured grids, a series of papers ad-
dressed, in the past, the practical implementation of these models and delineated achievements
or suggestions to improve their numerical behavior. In addition, practical problems about how to
approximate distances from solid walls or how to compute derivatives normal to the wall are
readily solvable on structured grids, with both precision and simplicity. In contrast, the relevant
approaches for unstructured grids with triangular elements are, by no means, trivial; they often
suffer from ambiguities or lack of precision during the treatment of the corresponding terms. For
instance, the grid stretching close to solid walls yields too skinny elements and, as a consequence,
standard discretization schemes become inaccurate. With the unstructured grids offering flexibility
in modeling complex domains, more systematic research in the direction of developing unstruc-
tured flow solvers has been undertaken in the last decade [3,4]. The ultimate goal is to remove the
relevant flaws, soften the lack of structure and exploit the indisputable advantages of easy
meshing.
Turbomachinery applications and the relevant complex flows are appropriate to support the
validation of turbulence models and codes. Thus, in this paper, the flow in a controlled-diffusion
compressor cascade [5] is analyzed, using six low-Reynolds turbulence models. They are incor-
porated in a finite volume, Navier–Stokes solver for unstructured grids with triangular elements.
All these turbulence models make use of the Boussinesq assumption; apart from the Spalart–
Allmaras one that solves a single differential equation, the rest utilize two transport equations for
the turbulent quantities (k–e or k–x). At off-design conditions, the compressor airfoil exhibits a
laminar leading edge separation bubble that triggers transition and an attached turbulent
boundary layer further downstream. We are concerned with the ability of the models to capture
the leading edge bubble and the subsequent development of the turbulent boundary layer under
adverse pressure gradient and streamline curvature. Several modifications to the standard models
are tested, as a means to overcome the unphysical turbulence growth close to the leading edge
stagnation point.
406 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

In the past, the same case has been studied numerically by other researchers, [1,2,6–8], for the
testing of various turbulence models. Many contributions to the study of this case are due to
Leschziner and his co-workers, [1,6,7]. They have concluded that conventional linear eddy-
viscosity models are not capable of resolving the leading edge separation bubble, unless a mech-
anism is introduced to suppress the excessive normal-strain-induced production of turbulence
energy. Thus, they worked out ways to bound the production terms [1] and, in a second step, they
implemented non-linear eddy-viscosity models [6,7]. Through the latter, an improved represen-
tation of the blade flow and a better resolution of the leading edge flow phenomena were obtained.
Tselepidakis and Kim [2] used the k–x model with modifications to the production term in order
to capture the separation bubble and quantitatively resolve the main flow characteristics. For the
same purpose, Kang et al. [8] employed a variant of the Launder–Sharma model with transition
modeling. Of course, the list of the aforementioned contributions is by no means exhaustive. Note
that a common feature of all of them is the use of structured grids, often with multiple blocks.
However, in the present work, unstructured grids with quasi-structured layers of triangular ele-
ments close to the solid walls have been used.

2. Governing equations

2.1. Mean flow equations

An approach that has often been pursued in the turbomachinery field is the quasi-3D analysis
of airfoil cascades through CFD tools. This requires 2D flow solvers where the effect of variable
streamtube thickness h ¼ hðx1 Þ has been properly taken into account. In this case, the corre-
sponding Favre-averaged compressible Navier–Stokes equations read
oðhWÞ oðhFi Þ
þ ¼S ð1Þ
ot oxi
T
where W ¼ ðq; qu1 ; qu2 ; Eeff Þ is the solution variable array and Fi ¼ Finv vis
i  Fi ði ¼ 1; 2Þ stand for
flux vectors. In non-dimensional form, their inviscid and viscous counterparts along with the
source term S and the viscous stresses, are given by
0 1 0 1 0 1
qui 0 0
B qui u1 þ peff di1 C B si1 C B dh C
Finv ¼ B C; Fvis ¼ B C; S ¼ hB ðp þ sh Þ dx1 C ð2Þ
i @ qui u2 þ peff di2 A i @ si2 A @ 0 A
oh
ðEeff þ peff Þui uj sij þ v oxi
0


leff 1 ouk u1 dh
sij ¼ 2 Sij  dij  dij ð3Þ
Re 3 oxk h dx1


4 leff u1 dh
sh ¼  S12 ð4Þ
3 Re h dx1
In order to effect closure, the state equation for perfect gases and a turbulence model have to be
considered, as well.
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 407

2.2. Turbulence models

In this work, six low-Reynolds turbulence models, have been programmed and tested. These
are listed below along with their abbreviations:
ðSAÞ Spalart–Allmaras one-equation model [9],
ðCHÞ Chien k–e two-equation model [10],
ðLSÞ Launder–Sharma k–e model [11],
ðWLÞ standard Wilcox k–x model [12],
ðBSLÞ blended k–e=k–x baseline model [13] and
ðSSTÞ shear stress transport variant of BSL [13].
All of them are based on the Boussinesq assumption and include no additional mechanism to
model transition. Apart from their standard formulations, modifications and refinements to the
aforesaid models have also been programmed and tested, as it will be discussed in a subsequent
paragraph. It is necessary to go through the models very briefly, in order to make the corre-
sponding expressions and constants as clear as possible.

2.2.1. The one-equation Spalart–Allmaras model


The SA model [9] solves a single differential equation for the quantity e l from which the tur-
bulent viscosity coefficient is derived. In the literature, this equation has been presented for
incompressible fluids. Herein, a variant for compressible fluids is presented which, considering
variable streamtube thickness, reads



oðhqelÞ o 1 1 o oe
l o ohe l
þ ðhqui e
lÞ ¼ ð1 þ cb2 Þ hðl þ e
lÞ  cb2 ðl þ e
lÞ þ SSA
ot oxi Re r oxi oxi oxi oxi
ð5Þ
"
2 #
1 e
l
lt ¼ fv1 e
l; SSA ¼ h cb1 q e
Sel  cw1 fw ð6Þ
Re n


e 1 e
l e
l v3 v
S ¼Xþ fv2 ; v¼ ; fv1 ¼ ; fv2 ¼ 1  ð7Þ
Re qj2 n2 l v3 þ c3v1 1 þ vfv1

1=6
1 þ c6w3 1 e
l
fw ¼ g 6 ; g ¼ r þ cw2 ðr6  rÞ; r¼ ð8Þ
g þ c6w3 Re e
S qj2 n2
where cv1 ¼ 7:1, cb1 ¼ 0:1355, cb2 ¼ 0:622, r ¼ 2=3, cw2 ¼ 0:3, cw3 ¼ 2, cw1 ¼ 3:239. Along the
solid walls, e
l is set to zero, whereas at the inlet, e
l is imposed in such a way that the inlet tur-
bulent viscosity equals a prescribed value.

2.2.2. Two-equation turbulence models


In vector form, all two-equation models can be cast as follows:
oðhWt Þ oðhFt;i Þ
þ ¼ St ð9Þ
ot oxi
408 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

where Wt ¼ ðqk; q/ÞT , / ¼ e or x, Ft;i ¼ Finv vis T


t;i  Ft;i and St ¼ hðSk ; S/ Þ . For the k–e and k–x
models, respectively, the turbulent viscosity is given by
qk 2 qk
lt ¼ cl fl Re ; lt ¼ Re ð10Þ
e x
k–e models: For the two k–e models used herein, the source terms are written as
e e2
Sk ¼ Pk  qe  D; Se ¼ ce1 f1 Pk  ce2 f2 q þ E ð11Þ
k k
where D and E are the low-Reynolds source terms. The production term in the kinetic energy
equation is given by


t 2 ouk
Pk ¼ sij  qk ð12Þ
3 oxk
where stij denotes the turbulent stresses (Eq. (3) with lt instead of leff ). In case of an eddy-viscosity
model this term is expressed by means of the turbulent energy generation term G as

2
lt 2 ouk oui 2 ouk u1 dh ouk
Pk ¼ G  qk ; G ¼ 2Sij  qk 2 ð13Þ
Re 3 oxk oxj 3 oxk h dx1 oxk
For the Chien (CH) model [10], the low-Reynolds terms are modeled by


l k l e nþ
D¼2 ; E ¼ 2 exp  ð14Þ
Re n2 Re n2 2
pffiffiffiffiffiffiffiffiffiffi
and fl ¼ 1  expð0:0115y þ Þ, f1 ¼ 1, f2 ¼ 1  0:22 exp½ðRet =6Þ2 , nþ ¼ Reðqn=lÞ sw =q, ce1 ¼
1:35, ce2 ¼ 1:80, cl ¼ 0:09, rk ¼ 1:0, re ¼ 1:3.
For the Launder–Sharma (LS) model [11]
pffiffiffi !2
2
l o k llt o2 Vs
D¼2 ; E¼2 ð15Þ
Re on qRe2 on2
 
and fl ¼ expð3:4=ð1 þ Ret =50Þ2 Þ, f1 ¼ 1, f2 ¼ 1  0:3 exp Re2t , Ret ¼ Reqk 2 =le, ce1 ¼ 1:44,
ce2 ¼ 1:92, cl ¼ 0:09, rk ¼ 1:0, re ¼ 1:3.
For both models, the wall boundary conditions are k ¼ 0; e ¼ 0. At the inflow, the following
2 3=2
relations are used: kin ¼ 1:5ðTu Vin Þ (Tu is available from measurements) and ein ¼ c3=4
l kin =L, where
L is set equal to 4% of the cascade pitch; the latter is a mixing length model estimate for fully
developed flows [14].
k–x models: For the standard Wilcox k–x model (WL), the source terms are given by [12]
x
Sk ¼ Pk  b
qkx; Sx ¼ c Pk  bqx2 ð16Þ
k
p ffiffiffiffi

where b ¼ 0:075, b
¼ 0:09, rx ¼ 0:5, rk ¼ 0:5, c ¼ b=b
 rx j2 = b
¼ 0:55:
For the blended k–e=k–x baseline (BSL) model [13] the source terms are
qRe 1 ok ox
Sk ¼ Pk  b
qkx; Sx ¼ c Pk  bqx2 þ 2qð1  F1 Þrx2 ð17Þ
lt x oxj oxj
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 409
( pffiffiffi ! )
  k 500l 4qrx2 k
F1 ¼ tanh arg14 ; arg1 ¼ min max
; 2
;
b xn Reqxn Ckx n2


1 ok ox
Ckx ¼ max 2qrx2 ; 1020
x oxj oxj
All the BSL model constants are computed as q ¼ F1 q1 þ ð1  F1 Þq2 , where q takes values in the
set q ¼ fb; rx ; rk ; cg and q1 ¼ f0:075; 0:5; 0:5; 0:55g, q2 ¼ f0:0828; 0:856; 1:0; 0:44g.
In the shear stress transport (SST) k–x model, introduced in Ref. [13] as a promising extension
of the BSL model, lt is given by
qk
lt ¼ ð18Þ
maxðx; XF2 =a1 Þ
   pffiffiffi 
where F2 ¼ tanh arg22 ; arg2 ¼ max 2 k =b
xn; 500l=Reqxn2 and the constants are taken as in
the q1 set with a new value for rk1 ðrk1 ¼ 0:85Þ.
The production term for the k–x models is limited [13], namely
Pk ¼ minðPk ; 20b
qkxÞ ð19Þ
The wall boundary conditions for the three k–x models are k ¼ 0 and xw ¼ 60l=Reqbn2w , where
nw is the distance of the first node off the wall [13]. Numerical tests have shown that, in un-
structured grids, the value of nw for any wall node P can be set equal to the mean of the heights
of the (two) triangles lying on the wall and containing this node as a vertex. At the inflow,
1=2
kin ¼ ð3=2ÞðTu Vin Þ2 and xin ¼ kin =cl L, with the mixing length L defined as above in the case of the
k–e models.

3. Numerical solution method

The governing equations are discretized on an unstructured grid with triangular elements using
a node-centered, finite-volume technique and solved through a time-marching scheme. At each
node P, the corresponding control volume CP is defined by successively connecting the midpoints
of the edges incident upon P with the barycentres of the surrounding elements (median dual
tessellation, Fig. 1). Both the mean flow and the turbulence equations are integrated over CP and
some area integrals are transformed into line integrals along its boundary oCP through the Green–
Gauss theorem. Thus
Z Z I Z Z
oðhWÞ
dx1 dx2 þ Fn ds ¼ S dx1 dx2 ð20Þ
CP ot oCP CP

where Fn ¼ ni hFi =jnj (n ¼ ðn1 ; n2 Þ is the boundary outward normal vector). Two different de-
compositions of oCP are used to account for the integration of the inviscid and viscous terms,
namely
[
ðoCP Þinv ¼ oCPQ þ ðoCP \ CÞ ð21Þ
Q2KN ðPÞ
410 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Fig. 1. Median dual tesselation of the unstructured grid.

[
ðoCP Þvis ¼ oCP;T þ ðoCP \ CÞ ð22Þ
T2KT ðPÞ

where C stands for the flow domain boundary. In view of the above, the line integrals take the
following discrete forms
I X
Finv
n ds ¼ Uinv ðWLPQ ; WR
PQ ; nPQ Þ ð23Þ
oCP Q2KN ðPÞ

I X
Fvis
n ds ¼ Uvis ðUG ; ðrUÞT ; nP;T Þ ð24Þ
oCP Q2KT ðPÞ

In the mean-flow equations, the inviscid numerical fluxes Uinv are computed by means of the
Roe flux-difference splitting scheme [15]. Second-order accuracy is achieved through the MUSCL-
T
extrapolation scheme applied to the primitive variables array U ¼ ðq; u1 ; u2 ; peff Þ , as follows:
1 1
UL ¼ UP þ WP PQ  ðrUÞP ; UR ¼ UQ  WQ PQ  ðrUÞQ ð25Þ
2 2
where W stands for the limiting function enforcing monotonicity to the solution and L, R denote
the ‘‘left’’ and ‘‘right’’ states in the corresponding Riemann problem [15]. According to Eq. (25),
the gradients of the primitive variables over the nodes should be computed. These are also needed
for the computation of the viscous stresses and the turbulent production term. At any node, a
Green–Gauss reconstruction [3] of the gradient is employed.
However, as suggested in Refs. [16,17], the use of very stretched unstructured grids calls for the
least squares approximation of the gradients involved in Eq. (25). The complementary use of
limiters (like, for instance, the 2D limiter of Barth and Jespersen [18]) is also necessary. From a
numerical point of view, the least squares scheme contributes to preserving the desired accuracy in
stretched grids, whereas the particular limiter is proved capable of overcoming non-monotone
situations, peaks and solution anomalies. In the past, the aforementioned requirements, as well as
the use of containment circle-dual tessellation of the domain [16] were verified in some other
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 411

applications of the present solver, where the grids exhibited extreme aspect ratios near the wall
[19]. However, for the grid used in the present study, which was generated using an excessive
number of nodes along the airfoil contour (see subsequent section), the cell aspect ratios close to
the walls were moderate and, consequently, it was found unnecessary to use either the least
squares approximation or any limiter.
In the turbulence equations, the inviscid numerical fluxes are computed by a first-order upwind
flux-splitting scheme, based on the local convective velocity normal to the cell boundary. For the
numerical calculation of the viscous numerical fluxes Uvis , the primitive variables U ¼ ðq; u1 ;
u2 ; peff ÞT are assumed to vary linearly within each triangle.
Local time stepping is used. The left-hand-side is derived by a first-order linearization of the
right-hand-side. The solution variables are updated in a loosely-coupled manner; first the mean-
flow equations (4  4 block system) are solved and then the turbulence model ones (2  2 system,
or a single equation for SA), both by the pointwise-implicit Gauss–Seidel scheme. Sub-iterations
(subscript k) are carried out within each time step, terminated whenever jdWk j=jdW0 j < 0:01,
(dWk is the solution of the linearized system of equations in the kth subiteration) or k ¼ kmax (here
kmax ¼ 15).

3.1. Practical implementation of the low-Reynolds models

The implementation of the low-Reynolds turbulence models in the context of a pointwise it-
erative solver requires the linearization of source terms, as a means to overcome numerical
problems caused by their stiffness in regions close to solid walls. In general, this linearization
(from time step n to n þ 1) reads
!n
nþ1 n oSðÞ
St ¼ St þ t
dWnþ1
t ð26Þ
oWt

where SðÞ
t is the part of the source term that yields a positive contribution to the diagonal term,
on the left-hand-side. For the k–e models, the linearization yields ðEþ ¼ maxðE; 0ÞÞ
!n n
ðÞ
oSke e=k  D=qk 1  D=qe
¼ ð27Þ
oWke ce2 f2 ðe=k Þ2 þ Eþ =qk 2ce2 f2 e=k þ Eþ =qe

For the k–x models, the linearization proposed by Menter [13] is used, namely
!n n
ðÞ
oSkx b
x 0
¼ ð28Þ
oWkx 0 2bx  jCkx j=qx

Finally, for the SA model, the linearization suggested in Ref. [9] is used, namely
 n
ðÞ n n
 oSSA =oqe l ¼ maxfðDSA  PSA Þ ; 0g þ maxfðD0SA  PSA 0
Þ ; 0gqel ð29Þ

where (see Eq. (6)) SSA ¼ qe


l ðDSA  PSA Þ, DSA ¼ DSA ðqe
l Þ and PSA ¼ PSA ðqe l Þ are the destruction
0 0
and production operators respectively and DSA ¼ oDSA =oðqe l Þ; PSA ¼ oPSA =oðqe l Þ.
412 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

4. Modifications to the two-equation models

In order to control the unphysical growth of turbulent kinetic energy near the leading edge
stagnation point, modifications to the conventional two-equation turbulence models have been
employed and tested. The first modification is related to the computation of the turbulence
production term in the k–e and k–x models and is based on the use of the local vorticity value.
The second one bounds the turbulent scale in the k–e and the k–x models (in all but the SST

Fig. 2. Different views of the computational grid. Upper: grid around the airfoil. Lower: close-up views of the grid close
to the leading (left) and the trailing (right) edges.
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 413

variant), as a means to enforce realizability to the turbulent stresses. The last two modifications
employ alternative ways to the computation of a spatially varying cl in the k–e models. The four
modifications are described below and assessed in the results section.

1. Two conceptually similar techniques to overcome the excessive turbulence production close to
the leading edge have been proposed by Kato–Launder [20] and Menter [13]. In the latter, the
production term is expressed in terms of the vorticity rather than the stress tensor, as follows:


2
ME lt 2 u1 dh ouk 2 ouk
Pk ¼ 2X  2  qk ð30Þ
Re h dx1 oxk 3 oxk
whereas in Ref. [20] the tensorial product of the vorticity and the mean rate of strain is used in
place of X2 . Here, we employed the Menter variant (denoted by ME) which was used along with
the k–e or k–x models.
2. As a means to enforce the realizability constraint to turbulent stresses, Durbin [21] imposed
computational bounds to the turbulent time scale Tt ¼ k=e. In the k–e model, Tt appears in both
the expressions of Se and lt , when cast in the forms Se ¼ ðce1 Pk  ce2 qeÞ=Tt þ E and lt ¼ cl qkTt ,
respectively. The expression proposed by Durbin for Tt is

Fig. 3. Conventional CH model. Effect of non-unit AVDR value.


414 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430
 
k 2 1
Tt ¼ min ; pffiffiffiffi ð31Þ
e 3cl G
This is readily generalized for use with the k–x model (since Tt ¼ k=e ¼ 1=ðcl xÞ), excluding the
SST variant, since a local bounding mechanism of lt already exists in this model by definition.
Through this modification, two-equation models become theoretically able to overcome the
unphysical production and growth of turbulent kinetic energy near the stagnation points. In
what follows, this variant will be denoted by DU (used with k–e or k–x models).
3. The third modification tested is the ‘‘variable cl ’’ version of the k–e models originated from the
Algebraic Reynolds Stress Model (ARSM) of Rodi [22], according to which cl is computed at
each grid node as

,
2
2 1  ec 1 Pk 1 Pk
cl ¼ 1 1  ec 1þ 1 ð32Þ
3 c1 c1 qe c1 qe

with c1 ¼ 1:5, ec ¼ 0:6. Though this model was originally proposed for non-equilibrium two-
dimensional thin shear layers, it has successfully been used herein in the expense of employing
upper and lower bounds for cl . This modification will be denoted by RO (used only with the
k–e model).

Fig. 4. Results of the CH model and its CH þ MR and CH þ RO variants.


D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 415

4. A generalized way to enforce realizability of the normal stresses, is the one proposed by Moore
and Moore et al. [14]. According to this, cl is replaced by the following expression
1
cl;r ¼ ð33Þ
ea e a Þ1=a
Ao þ As ð S þ AX X
where Ao ¼ 2:74, As ¼ 1:9, AX ¼ 1 and a ¼ 2:4. This substitution is realized in practice by
multiplying the expression of the lt coefficient definition by fn ¼ minð1; cl;r =cl fl Þ, in order to
allow for the activation of the model’s damping function fl near the low-Reynolds regions. This
modification will be denoted by MR (used only with the k–e model).

5. Results and discussion

The problem selected for the comparative study of the aforementioned turbulence models is
concerned with the flow in highly-loaded controlled-diffusion compressor cascade. In the past, this
case has been investigated experimentally by Elazar and Shreeve [5], at design and off-design
flow conditions. Here, emphasis is laid on the investigation and interpretation of the predictive

Fig. 5. Results of the CH model and its CH þ DU and CH þ ME variants.


416 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

capabilities of the turbulence models, without and with the aforesaid modifications. The flow
conditions are: inlet Mach number M1 ¼ 0:25, Reynolds number based on chord Rec ¼ 700; 000,
inlet turbulence intensity Tu ¼ 1:5%. The case with inlet flow angle a1 ¼ 46° that corresponds to
off-design operations, will be presented first and in all detail. Then, results for a milder a1 value
ða1 ¼ 43:4°Þ and near-design conditions ða1 ¼ 40°Þ will be presented. As reported in Ref. [5], in the
a1 ¼ 46° case, a laminar separation bubble appears over the suction side and close to the leading
edge. The early transition caused by this bubble yields a turbulent flow developing under strong
adverse pressure gradient along the major part of the blade suction side. The turbulence field at
the leading edge affects the boundary layer development further downstream. It should be pointed
out that, despite the large incidence, the airfoil is designed to maintain the attached flow all over
the suction side, so that the a1 ¼ 46° case is an appealing problem for the validation of turbulence
models.
The blade airfoil coordinates are given in Ref. [5]. The blade chord is 12.73 cm, the cascade
pitch 7.72 cm and the stagger angle 14:4°. The streamtube thickness along the axial direction is not
uniform, as discussed in a subsequent paragraph.
A single unstructured grid with triangular elements was generated and used. Since low-Rey-
nolds turbulence models call for an intensive grid resolution near the wall, a ‘‘structured-like’’ grid
of triangular elements in the vicinity of the blade contour was employed. This consists of 16 O-
type layers which, considering that 2000 nodes were placed along the blade contour, yield about
32,000 nodes in the ‘‘boundary layer’’ region. The remaining part of the computational domain
was filled with triangular elements in a fully unstructured manner using an in-house mesh gen-
erator. The total grid size is 43,321 nodes/84,542 triangles (Fig. 2). The clustering of grid nodes
close to the solid wall is such that the non-dimensional distance nþ of the first nodes off the wall
remained below 1. Though the relevant study is omitted in this paper, it should be pointed out
that the so-obtained results are grid independent. No particular effort was made to minimize the
total number of nodes required for grid independent solutions.

Fig. 6. CH model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0157).
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 417

Figs. 3–26 correspond to the largest incidence angle, a1 ¼ 46°. Some of them consist of several
plots illustrating the parallel to the wall velocity and turbulence intensity profiles at three axial
positions over the blade suction side, as well as pressure coefficient, displacement thickness and
skin-friction coefficient distributions along the airfoil contour. Using these plots, one may com-
pare all but the friction coefficient distributions with the measured ones [5], identify the existence
and extent of the leading edge bubble and diagnose tendency to separation. In some other figures,
streamlines and turbulence kinetic energy contours are plotted in the vicinity of the leading edge,
as a means to shed light into the formation of the separation bubble and the turbulence level in
that region.
In contrast to previous studies of the same cascade [7,8] which did not account for the non-unit
AVDR value mentioned in Ref. [5], here the streamtube thickness h ¼ hðx1 Þ was considered to

Fig. 7. CH þ MR model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0110).

Fig. 8. CH þ DU model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0137).
418 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

vary in the axial direction x1 . Practically, h remained constant upstream and downstream of the
blades and varied linearly between the leading and trailing edges, so that AVDR ¼ 1=1:025.
Motivated by the reviewers’ comments, a study was undertaken in order to investigate the role
of the non-constant streamtube thickness distribution along the axial direction. For this pur-
pose, two computations were carried out, both using the standard CH model. In the first, the
AVDR ¼ 1=1:025 value was used, whereas in the second AVDR ¼ 1. As shown in Fig. 3, results
along the front part of the airfoil are identical. However, differences can be observed along the
rear part of the suction side. With AVDR ¼ 1, the velocity profile or displacement thickness
plots at the last station, and especially the friction coefficient distribution along a considerable
percentage of the chord, indicate a separated turbulent flow. Since separation in this area
was never confirmed experimentally, this study indicates that the correct AVDR value should
be employed in the computations. The AVDR ¼ 1 computation yields a thicker boundary
layer along the second half of the blade and different transversal profiles at the trailing edge
station.
Figs. 4 and 5 summarize results obtained using the CH model and its variants. For the purpose
of comparison, the CH model results are repeated in both figures. In its conventional form, the
CH model predicts a quite small separation bubble close to the leading edge. At the first station
(5.9% of chord downstream of the leading edge), the computed velocity profile does not match the
measured one, as this station lies close to or in the bubble. Also, the velocity at the boundary layer
edge is slightly overestimated. However, further downstream, the CH model yields satisfactory
predictions; it should be noted that the predicted momentum thickness distribution along the
suction side almost coincides with measurements. It is known (see for example Ref. [1]) that the
excessive production of turbulent kinetic energy k close to the leading edge is the main reason for
local discrepancies. In Fig. 6, the unphysical concentration of iso-k contours near the leading edge

Fig. 9. Typical convergence of the (four) mean flow and the (two) turbulence model equations. The CH þ MR model
has been used.
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 419

Fig. 10. Results of the LS model and its LS þ MR and LS þ RO variants.

is visible. In the presence of high k values, the conventional CH model predicts an extremely small
bubble that can be hardly identified when drawing streamlines (Fig. 6).
In order to eliminate the excessive production of k at the leading edge, the CH model was used
along with four modifications, as exposed previously in this paper. Results obtained using the
CH þ MR and CH þ RO variants, i.e. by employing the Moore and Moore [14] and the Rodi
[22] modifications respectively, are demonstrated in Fig. 4. Both exhibit similar behavior, as they
give rise to a more lengthy separation bubble of considerable thickness compared to the con-
ventional CH model. Also, the velocity profile at the first station is much closer to the measured
one, due to the improved flow simulation at the leading edge region. In the same region, a local
improvement in the pressure coefficient distribution with respect to measurements can be re-
ported. Despite the different flow patterns at the leading edge, these modifications seem to affect
negligibly the flow development along the rear part of the blade, where the displacement thick-
ness distribution remains practically the same to that computed by the CH model. Fig. 7 illus-
trates the iso-k contours in the vicinity of the leading edge as computed by the CH þ MR variant
and has to be compared with Fig. 6 to enlight the effect of this modification. A similar leading
edge bubble and k field have been obtained using the CH þ RO variant, so these figures can
be omitted.
420 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Fig. 11. Results of the LS model and its LS þ DU and LS þ ME variants.

Fig. 12. LS model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0100).
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 421

Fig. 13. LS þ RO model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0060).

Results obtained using the other two variants, namely CH þ DU and CH þ ME, based on the
Durbin [21] and Menter [13] modifications respectively, are shown in Fig. 5. The friction coeffi-
cient distributions indicate that they both tend to produce a bubble of noticeable length and
thickness. The flow patterns and the iso-k contours at the leading edge using CH þ DU and
CH þ ME are quite similar; plots corresponding to CH þ DU are presented in Fig. 8. The velocity
kink at the first station and the pressure coefficient in the same region are better captured.
However, in contrast to the previously tested variants, the flow downstream has been affected by
the upstream flow, as it can be seen by examining the longitudinal velocity profiles at the third
station. These profiles, predicted by the CH þ DU and CH þ ME variants, are slightly thicker
and lead to underprediction of the displacement thickness distribution.
We recall that all of the results are grid independent (as exposed elsewhere in this text) and fully
converged. An typical convergence history is shown in Fig. 9 for the CH þ MR model.

Fig. 14. LS þ ME model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0096).
422 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Fig. 15. Results of the WL model and its WL þ DU and WL þ ME variants.

Fig. 16. WL model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0090).
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 423

Fig. 17. WL þ ME model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0068).

Fig. 18. WL þ DU model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0059).

Having in mind the predictive capabilities of the CH model and its variants, we may easily go
through the Launder–Sharma k–e model (LS). While the CH model captured a very small bubble,
the LS model fails to capture any bubble. It seems that, with the LS model, the concentration of
the iso-k contours close to the leading edge (Fig. 12) is pronounced compared to that of the CH
model (Fig. 6). By employing the same four modifications to the LS model, the problem is al-
leviated and the local flow patterns come closer to measurements. Fig. 10 (for the LS, LS þ MR
and LS þ RO) and Fig. 11 (for the LS, LS þ DU and LS þ ME) summarize the relevant nu-
merical predictions. The lengthy bubbles predicted by the LS þ MR and LS þ RO models seem to
affect the flow downstream. In fact, using these models, the flow at the third station is too close
to separation, as it becomes evident from the friction coefficient plots. A second effect is that
they overestimate the displacement thickness values with respect to the already overestimated
424 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Fig. 19. Results of the BSL model and its BSL þ ME and BSL þ DU variants.

Fig. 20. BSL model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0081).
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 425

Fig. 21. BSL þ ME model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0103).

Fig. 22. BSL þ DU model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0092).

distribution that the conventional LS model produced. In contrast to the MR and RO modifi-
cations, the LS þ ME and LS þ DU variants produce smaller bubbles without considerably
affecting the velocity profile downstream (third station). The latter is in good agreement
with measurements and with no tendency to separate. An immediate consequence is that
the displacement thickness distribution is underestimated compared to LS results and comes
closer to the measured ones. The velocity and turbulence kinetic energy fields are shown in Figs.
12–14.
Fig. 15 compares results obtained using the standard k–x model (WL) and its variants
(WL þ ME, WL þ DU). The conventional model produces a small bubble at the leading edge
with discrepancies in the velocity profile at the first station. The WL þ ME and WL þ DU
variants predict bubbles of increased size and this reflects on the velocity profiles at the first
426 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

Fig. 23. Results of the SST model and its SST þ ME variant.

Fig. 24. SST model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0064).
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 427

Fig. 25. SST þ ME model. Left: streamlines. Right: 20 iso-k contours (min–max values: 0–0.0061).

Fig. 26. Results of the SA model and the predicted leading edge bubble.
428 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

station which come closer to the measured one. Differences in the bubble size can be seen in the
pressure or friction coefficient plots. Downstream of the first station, the three models yield
quite similar results, the WL þ DU however overpredicts the displacement thickness. In general,
either without or with modifications, the WL model results to a slight separation towards the
trailing edge. The corresponding velocity and turbulence kinetic energy fields are shown in Figs.
16–18.
Fig. 19 presents results from a similar study, where the BSL model and its variants have been
used. With the BSL model(s), the flow in the rear part of the suction side is that far from sepa-
ration that the velocity profile at the third station is thicker and deviates considerably from the
measured one. As a consequence, the displacement thickness distribution over the suction side is
underpredicted. At the leading edge, the bubble predicted using the BSL model is very small but
its size increases substantially with BSL þ ME or BSL þ DU models (see Figs. 20–22).
Results obtained using the SST and SST þ ME models are shown in Fig. 23. The SST model
predicts a rather long bubble which increases if the Menter modification is used. As shown in Figs.
24 and 25, the SST þ ME model predicts a thicker bubble which is located further downstream
compared to that predicted by SST. With both variants, the flow at the rear part of the suction
side is on the verge of separation and the velocity profile at the third station is captured very
satisfactorily.
Results obtained using the SA model are illustrated in Fig. 26. Since the SA model does not
take into account the incoming turbulence (which, in the present case, is relatively low, about
1.4%), its results are sensitive only to the inlet boundary condition for lt . Here, a value of
lt;in ¼ 50l has been used. With this value, the SA model predicts a laminar separation bubble of
reasonable size and the flow is not separated downstream. The predicted velocity profiles at the
three stations are in satisfactory agreement with the experimental data.
For the sake of completeness, Figs. 27 and 28 present the pressure coefficient distributions for
two more operating points, namely the a1 ¼ 40° (near-design point) and the a1 ¼ 43:2° case. The
CH þ MR and the SST þ ME models have been used and their results are in satisfactory
agreement with measurements.

Fig. 27. Pressure coefficient comparison of the CH þ MR and SST þ ME results against experimental data for the 40°
inlet angle case.
D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430 429

Fig. 28. Pressure coefficient comparison of the CH þ MR and SST þ ME results against experimental data for the
43.4° inlet angle case.

6. Conclusions

Six popular low-Reynolds number turbulence models have been employed in a time-marching
Navier–Stokes solver. Computations were carried out using an unstructured grid with structured-
like layers in the vicinity of the solid walls. This enabled the balanced use of high spatial resolution
in the boundary layer along with lower storage requirements elsewhere.
The turbulence models were tested on the prediction of the flow in a controlled-diffusion
compressor cascade, at off-design conditions. This case is characterized by the formation of a
laminar separation bubble at the leading edge, while the flow remains fully attached along the
suction side, despite the strong adverse pressure gradient.
As expected, the conventional k–e and k–x models fail to capture the laminar bubble due to the
unphysical production of turbulence kinetic energy near the stagnation point. For instance, the
Launder–Sharma model predicts no bubble at all, while the Chien model results to a small bubble
which is by no means laminar. Among the goals of this paper was to evaluate the performance of
four modifications that have been proposed in the literature to alleviate this problem. Modifi-
cations based on variable cl coefficient and/or different expressions for the production term sat-
isfactorily predicted the presence of the leading edge bubble with both k–e models. None of the k–e
models leads to flow separation at the rear part of the suction side.
The Wilcox k–x model yields a separated flow over the major part of the suction side, re-
gardless of the appearance and size of the leading edge bubble. The latter is a consequence of the
use of the previous modifications. The baseline k–e=k–x model, with or without modifications,
yields fully attached flow in the rear part of the suction side. At the leading edge, the bubble
predicted using this model is very small but its size increases substantially by eliminating the
excessive production of k. The SST model proposed by Menter predicts a rather long bubble
which becomes thicker and moves downstream by using a modified expression for k production.
The one-equation SA model does not suffer from unphysical production of turbulence near the
stagnation points but its inability to account for the incoming turbulence in internal aerodynamics
is troublesome. With carefully selected inlet boundary conditions the use of the SA model may
lead to satisfactory predictions.
430 D.G. Koubogiannis et al. / Computers & Fluids 32 (2003) 403–430

References

[1] Chen WL, Lien FS, Leschziner MA. Computational modelling of turbulent flow in a turbomachine passage with
low-re two-equation models. In: Wagner S, Hirschel EH, Periaux J, Riva P, editors. Proceedings of the Second
European Computational Fluid Dynamics Conference, ECCOMAS. Stuttgart, Germany: Wiley; 1994. p. 517–24.
[2] Tselepidakis DP, Kim SE. Modeling and prediction of the laminar leading-edge separation and transition in a
blade cascade flow. ASME Paper 96-GT-411, 1996.
[3] Barth TJ. Aspects of unstructured grids and finite-volume solvers for the Euler and Navier–Stokes equations.
AGARD Report 787. Special Course on Unstructured Grid Methods for Advection Dominated Flows, 1992.
[4] Farhat C, Lanteri S. Simulation of compressible viscous flows on a variety of MPPs: computational algorithms for
unstructured dynamic meshes and performance results. Comp Meth Appl Mech Engng 1994;119:35–60.
[5] Elazar Y, Shreeve RP. Viscous flow in a controlled diffusion compressor cascade with increasing incidence. Trans
ASME J Turbomachinery 1990;112:256–66.
[6] Chen WL, Lien FS, Leschziner MA. Computational modelling of cascade-blade flow with linear and non-linear
low-re eddy-viscosity models. In: AGARD-CP-571. Loss Mechanisms and Unsteady Flows in Turbomachines.
1995.
[7] Chen WL, Lien FS, Leschziner MA. Non-linear Eddy-viscosity modelling of transitional boundary layers pertinent
to turbomachine aerodynamics. Int J Heat Fluid Flow 1998;19:297–306.
[8] Kang SH, Lee JS, Choi MR, Kim KY. Numerical calculations of the turbulent flow through a controlled diffusion
cascade. Trans ASME J Turbomachinery 1995;117:223–30.
[9] Spalart P, Allmaras S. A one-equation turbulence model for aerodynamic flows. AIAA Paper 92-0439, 1992.
[10] Chien KY. Predictions of channel and boundary-layer flows with a low-Reynolds-number turbulence model. AIAA
J 1992;20(1):33–8.
[11] Launder BE, Sharma BI. Application of the energy disspation model of turbulence to the calculation of flow near a
spinning disc. Lett Heat Mass Transf 1974;1(2):131–8.
[12] Wilcox DC. Turbulence modeling for CFD. DCW Industries, 1993.
[13] Menter FR. Zonal two equation k–x turbulence models for aerodynamic flows. AIAA Paper 93-2906, 1993.
[14] Moore JG, Moore J. Realizability in turbulence modelling for turbomachinery CFD. ASME Paper 99-GT-24,
1999.
[15] Roe P. Approximate Riemann solvers, parameter vectors, and difference schemes. J Comp Phys 1981;43:357–71.
[16] Barth TJ. Numerical aspects of computing viscous high Reynolds number flows on unstructured meshes. AIAA
Paper 91-0721, 1991.
[17] Anderson WK, Bonhaus DL. An implicit upwind algorithm for computing turbulent flows on unstructured grids.
Comput Fluids 1994;23(1):1–21.
[18] Barth TJ, Jespersen D. The design and application of upwind schemes on unstructured meshes. AIAA Paper 89-
0366, 1989.
[19] Koubogiannis DG, Giannakoglou KC. Implementation and assessment of low-Reynolds turbulence models for
airfoil flows on unstructured grids. In: CD-Rom. Proceedings of the European Congress on Computational
Methods in Applied Sciences and Engineering Fluid Dynamics Conference, ECCOMAS 2000. Barcelona, Spain,
2000.
[20] Kato M, Launder BE. The modelling of turbulent flow around stationary and vibrating square cylinders. Proc 9th
Turbulent Shear Flows, Kyoto, Japan, 1993. p. 456.
[21] Durbin PA. On the k– stagnation point anomaly. Int J Heat Fluid Flow 1996;17:89–90.
[22] Rodi W. A new algebraic relation for calculating the Reynolds stresses. ZAMM 1976;56:219–21.

You might also like