Advanced Turbulence Modelling of Separated Flow in A Diffuser

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Flow, Turbulence and Combustion 63: 81–112, 1999.

81
© 2000 Kluwer Academic Publishers. Printed in the Netherlands.

Advanced Turbulence Modelling of Separated


Flow in a Diffuser

D.D. APSLEY
Department of Civil & Structural Engineering, UMIST, Manchester M60 1QD, U.K.

M.A. LESCHZINER
Department of Engineering, Queen Mary and Westfield College, University of London,
London E1 4NS, U.K.

Abstract. The paper describes an investigation into the predictive performance of linear and non-
linear eddy-viscosity models and differential stress-transport closures for separated flow in a nom-
inally two-dimensional, asymmetric diffuser. The test case forms part of a broader collaborative
exercise between academic and industrial partners. It is demonstrated that advanced turbulence mod-
els using strain-dependent coefficients and anisotropy-resolving closure offer tangible advantages
in predictive capability, although the quality of their performance can vary significantly, depend-
ing on the details of closure approximations adopted. Certain features of the flow defy resolution
by any of the closures investigated. In particular, no model resolves correctly the flow near the
diffuser’s inclined wall immediately downstream of the inlet corner, which may reflect the pres-
ence of a “flapping” motion associated with a highly-localised process of unsteady separation and
reattachment.

Key words: turbulence modelling, separated flow, non-linear eddy-viscosity models, differential
stress models.

1. Introduction
Half a century of research in turbulence modelling has spawned numerous closure
models arising from many permutations of intuitive concepts, rational principles
and calibration practices. The largest, oldest and most widely-used class of models
is that combining the eddy-viscosity concept with the linear stress-strain relation-
ship and one or two transport equations for respective surrogates of the turbulent
velocity and length scales. The recognition that this type of model is afflicted with
serious weaknesses, especially in the presence of high curvature, swirl, separation,
stagnation and body forces, has led the turbulence-research community to search
for superior models which are based on a firm physical foundation.
The three principal routes taken over the past few years have been based on
non-linear stress-strain relations [7, 8, 11], transport equations for the stresses
[12, 16, 18, 28] and scalar fluxes and, most recently, on equations for turbulence-
structure parameters [26]. Although there is widespread recognition of the funda-
82 D.D. APSLEY AND M.A. LESCHZINER

mental strengths and superior predictive potential of such models, CFD practition-
ers in industry and vendors of commercial CFD software continue to rely heavily
on linear two-equation models, and have not, on the whole, taken more than tent-
ative steps towards adopting more advanced forms. Indeed, there has been a recent
trend in the aerospace industry (especially in the USA) towards even simpler one-
equation models [2, 27] which rely, intrinsically, on an empirical prescription of a
turbulent length scale. In principle, such models are suited to flows which are close
to equilibrium and in which a single shear stress dominates the linkage between
the mean flow and turbulence.
The reluctance of CFD practitioners to adopt advanced models is rooted in a
number of practical considerations. Greater mathematical and coding complexity,
lower numerical robustness, more stringent grid-quality requirements and more
extensive reliance on boundary conditions for turbulence quantities are some of
the reasons often put forward, and most of these are, indeed, based in reality.
Perhaps the most powerful and potentially damaging argument arises, however,
from the perception that there is no convincing evidence that advanced models per-
form consistently better than carefully-tuned simple models over a broad range of
conditions. This perception is heightened by the outcome of a number of influential
workshops (e.g. [6, 15]) which involved, for prescribed test flows, comparisons of
computational solutions contributed by a number of participants using their own
codes, grids, numerical practices and model implementations. Clearly, within such
loosely-controlled exercises, there is extensive potential for variability in predictive
performance, which may be unrelated to the intrinsic properties of the turbulence
models supposedly being investigated. Thus, occasionally, if not usually, the only
conclusion that can safely be extracted from an extensive exercise of this type is
that it is inconclusive and that an interactive, more carefully-controlled initiative
would be required to remove inexplicable features which had come to light in the
comparisons undertaken. In most circumstances, any follow-up activity is limited
in scope and not very fruitful. The overall damaging consequence is an even greater
level of scepticism and uncertainty than before, and a further strengthening in the
reluctance of practitioners to adopt advanced modelling practices.
The urgent need perceived by industry to gain a much clearer view of the relative
performance of different turbulence models in challenging conditions akin to those
encountered in practice led, in 1997, to the formation of a consortium of industrial
partners (BAE SYSTEMS, Aircraft Research Association, Rolls-Royce, Defence
Evaluation and Research Agency) and academic groups (UMIST, Loughborough
University of Technology) with the goal of undertaking systematic, collaborative,
computational investigations of a selection of well-documented flows spanning
a broad range of flow-physical and geometric features. The consortium, work-
ing within the VoTMATA (Validation of Turbulence Models for Aerospace and
Turbomachinery Applications) project, seeks to provide an objective assessment of
existing turbulence models, in various classes, for flows involving strong curvature,
separation from smooth surfaces, 3-d strain, strong vorticity, impingement and
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 83

shocks. The first three test cases – a 2-d asymmetric diffuser, a transonic axisym-
metric bump and a 3-d wing-body junction – are comparatively simple in terms
of geometry, and are characterised by detailed and accurate experimental data. A
large number of turbulence models have been used to calculate each test case under
strictly controlled numerical conditions.
This paper focuses on the first of the “fundamental” test cases: the asymmetric
plane diffuser of Obi et al. [24]. This flow is of considerable interest to turbulence
modellers, and has featured as one of three test cases in a recent IAHR/ERCOFTAC
Workshop [15]. Besides the original computations of Obi et al. with high-Re k–ε
and differential stress models, it has also been computed by Durbin [9] with the k–
ε–v 2 model which was reported to perform well. The attraction of this test case lies
in its geometric simplicity, its virtually two-dimensional state, the availability of
two sets of closely matching experimental data for mean velocity, turbulence and
surface coefficients, obtained in two separate facilities [4, 24], and the existence
of LES results [10] which further support the validity of the experimental data.
The test case presents a challenging example of separation from a plane surface
in an adverse pressure gradient. The experimental configuration is described in
Section 2.
For geometries where separation is not fixed by sharp corners, determining the
separation line depends on modelling the correct response of the shear stress and,
to some degree also the normal stresses, to deceleration; that is, of predicting the
correct turbulent time scale. Beyond the separation point, the flow loses much of
the direction-constraining influence of the boundary, and all components of the
Reynolds-stress tensor become dynamically significant. At the same time, the rate
of recovery and, hence, the reattachment point, depend on the magnitude of the
shear stress in a (curved) free shear layer. The challenges for turbulence models in
such a flow are, therefore, (i) to predict the correct turbulent time scale in strongly
accelerated wall and free shear layers; and (ii) to predict the correct anisotropy of
the Reynolds-stress tensor.
The turbulence models applied to this test case have been somewhat loosely
categorised into four classes. The minimum degree of complexity is represented
by two-equation, constant-coefficient, linear eddy-viscosity models (EVMs). A
second class of models is formed from the hybrid k–ω/k–ε models of Menter
[23], including the shear-stress transport (SST) scheme. The last two classes
are populated by anisotropy-resolving closures: non-linear eddy-viscosity models
(NLEVMs) and differential stress models (DSMs). Throughout, there are various
choices of turbulence scalars (k–ε, k–ω, k–g) and wall boundary conditions. The
turbulence models and a theoretical assessment of model differences and capab-
ilities are given in Section 3. Section 4 contains a detailed comparison of model
performance against the experimental data. Section 5 attempts to summarise what
we have learnt.
84 D.D. APSLEY AND M.A. LESCHZINER

Figure 1. Test-case geometry.

Figure 2. Finite-volume mesh used for low-Re calculations; (every other grid line omitted for
clarity).

2. Description of the Test Case


The test case involves separation from a plane wall and subsequent reattachment
in the downstream duct of a diffuser. The geometry is shown in Figure 1. The
diffusing section has a length 21H , where H is the inlet channel height, and overall
expansion ratio 4.7. The Reynolds number, based on an upstream reference velocity
and H, is 21200. The flow was observed to separate from the angled wall at x/H ≈
7.5 and to reattach at x/H ≈ 29, where x is measured from the start of the diffuser.
In Obi et al.’s experiment, profiles of U, V, u2 , v 2 and uv were measured by two-
component laser-doppler anemometry (LDA) at stations between x/H = 3.2 and
x/H = 25.2. Pressure data were obtained from the plane wall.
The inlet duct was of length 115H , sufficient to ensure a fully-developed
turbulent channel flow. Inlet boundary conditions (see below) were based on exper-
imental measurements of U and u2 by single-wire hot-wire anemometry (HWA)
at x/H = −11. The aspect ratio was 1:35 at inflow and 1:7.4 at outflow. The
latter is possibly insufficient to ensure strict two-dimensionality. However, a sub-
sequent repeat of the experiment by Buice and Eaton [4] at Stanford University,
with careful attention to eliminating side-wall effects, supported Obi et al.’s data in
the measurement region covered. Buice and Eaton measured velocities using hot-
wire probes and were unable to make measurements in the recirculating flow. They
did, however, extend the coverage of flow development downstream, and added
surface-pressure and skin-friction coefficients to the data set.
The computational domain extended from x/H = −11 to x/H = 60, em-
ploying a finite-volume mesh (Figure 2) of 292 × 96 control volumes (minimum
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 85

cell height 0.001H ) for low-Re calculations and 292 × 56 control volumes for
high-Re calculations with wall functions. In the latter case, low-Re-grid cells at
less than 5% channel height from each wall were amalgamated, without changing
the interior disposition of control volumes. The near-wall grid sizes were chosen
to obtain a y + value at cell-centres (based on a log-law fit to the inflow velocity
profile) of 1 for low-Re and 25 for high-Re calculations.
Inflow profiles were determined for each turbulence model by a preliminary
computation (with the same y grid) of fully-developed channel flow, with the
same mass flow rate as the experimental data. The remaining boundary conditions
were zero longitudinal gradients for all transport variables at outflow and no-slip
conditions at the diffuser walls.

3. Turbulence Modelling
3.1. L INEAR EDDY- VISCOSITY MODELS

The simplest form of turbulence closure that preserves the symmetries and tensorial
nature of the Reynolds stresses is a linear eddy-viscosity model (EVM), which
assumes a simple proportionality between deviatoric stress and mean strain:
 
2 1
ui uj − kδij = −2vt Sij − δij Skk , (1)
3 3
where
 
1 ∂Ui ∂Uj
Sij = + .
2 ∂xj ∂xi

The turbulent kinetic energy k ≡ (1/2)ui ui must be specified – usually by solving


its own transport equation.
On (primarily) dimensional grounds, vt may be written as a product

vt = u0 l0 = u20 t0 , (2)

where u0 is a magnitude of the cross-stream turbulent fluctuations, l0 is a “mixing”


length and t0 is a turbulence time scale. Typically, u0 ∝ k 1/2 and t0 is propor-
tional to the eddy turnover time k/ε, where ε is the rate of dissipation of k. ε
may be modelled by specifying a dissipation length lε = u30 /ε, algebraically,
in terms of wall-normal distance. In geometrically complex flows, however, it is
common practice to solve a second transport equation, either for ε itself, or for
some related dimensional variable (e.g. ω). The majority of (linear and non-linear)
eddy-viscosity models in use today require the solution of transport equations for
two turbulent scalars.
Models investigated herein relate the eddy viscosity to different turbulence-
velocity and length-scale-determining variables. There is little consensus, but keen
86 D.D. APSLEY AND M.A. LESCHZINER

partisanship, amongst turbulence modellers regarding the “optimal” turbulence


transport variables. Alternative proposals examined here are as follows:
k–ε models: vt = Cµ fµ k 2 /ε, t0 ∼ k/ε,
k–ω models: vt = fµ k/ω, t0 ∼ ω−1 ,
k–g models: vt = Cµ fµ kg 2 , t0 ∼ g 2 . (3)
In the above, fµ represents a viscous damping term. At a solid boundary, theory
predicts the following asymptotic behaviour as wall-normal distance y → 0:
2vk
k ∝ y2, ε∼ ∼ constant,
y2
2v y
ω∼ → ∞, g∼ → 0. (4)
Cµ y 2 (2ν)1/2
Note that some k–ε models, notably that of Launder and Sharma [19], use the
isotropic part of the dissipation rate, ε = εabs − 2v(∂k 1/2 /∂xi )2 (usually written
as ε̃), in which case the boundary condition becomes ε → 0 (y → 0). The wall-
boundary condition for ω clearly requires a “numerical fix”; the most commonly
accepted practice [23] specifies a finite value of ω on the boundary. The use of
1/2
g ∝ t0 is designed for the numerical convenience of having a variable which
varies linearly with distance from the wall; apparently, this is ignored for the other
transport variable, k.
For k–ε, k–ω and k–g models the turbulent transport equations can be written,
for incompressible flow, as
  
Dk vt
= ∇· v+ ∇k + P − εabs
Dt σk
  
Dγ vt γ
= ∇· v+ ∇γ + (Cγ 1 P − Cγ 2 ε) + Sγ , (5)
Dt σγ k
where γ represents any of ε, ω, g, etc.,
∂Ui
P ≡ −ui uj (6)
∂xj
is the rate of production of turbulent kinetic energy, and ε can be determined by
combining two eddy-viscosity relations in (3). Cγ 1 , Cγ 2 and Sγ may include the
effects of molecular viscosity.
Writing
γ = Aε m k n , (7)
then the second transport equation can always be transformed into one for ε:
  
Dε vt ε
=∇· v+ ∇ε + (Cε1 P − Cε2 ε) + Sε , (8)
Dt σε k
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 87
Table I. Coefficients in the length-scale-determining equation.

γ A m n Cγ 1 Cγ 2 Cε1 Cε2

Launder and Sharma [19] ε 1 1 0 1.44 1.92 1.44 1.92


Wilcox [30] ω −1
Cµ 1 −1 0.56 0.83 1.56 1.83
Kalitzin et al. [17] g 1 –1/2 1/2 –0.28 –0.42 1.56 1.83

where
Cγ 1 − n Cγ 2 − n
Cε1 = , Cε2 = (9)
m m
and (if ε = εabs )
     
ε 1 1 nε 1 1
Sε = Sγ + ∇ · − vt ∇ε + ∇· − vt ∇k
γm σγ σε mk σγ σk
  "   2  2 #
ε vt ∇ε ∇k 2 ∇ε ∇k
+ v+ m +n −m −n . (10)
m σγ ε k ε k
Clearly, all two-equation turbulence models of this type can be interconverted. The
corresponding values of Cε1 and Cε2 deduced from the coefficients of k–ω and k–g
models are given in Table I.
The additional source term consists of parts due to different turbulent Prandtl
numbers and to cross-diffusion. Both stem from the original diffusion model and do
not contain mean-velocity gradients explicitly. In general, it is difficult to establish
the precise effect of the cross-diffusion terms which may be of either sign. More
clear-cut, however, is the effect of the difference Cε2 –Cε1 , particularly as regards
the response to adverse pressure gradients and the prediction of separation: the
smaller the difference between these constants, the greater is ε and hence the less
is the turbulence intensity. Experience shows that the shear stress is extremely sens-
itive to this difference, even though both k–ε and k–ω models satisfy the log-law
in equilibrium flows (for an appropriate choice of σε and σω ). Although the role of
the cross-diffusion terms is not easily analysed, the values of Cε1 and Cε2 in Table I
help to explain why k–ω models (with their standard calibration) are more likely
than k–ε models to predict separation in adverse pressure gradients and, conversely,
why k–ω models tend to over-predict reattachment length in bluff-body separated
flows.
The eddy-turnover time scale is t0 ∼ k/ε. Mean-velocity gradients give rise to
a second (“shear”) time scale, T0 ∼ kSk−1 . Thus, from (1) and (2),

ui uj /u20 ∼ Cµ t0 /T0 . (11)


In order that u20
should continue to reflect the scale of turbulence energy, i.e. that
the LHS of (11) should remain of similar size for large shear, it is required that
88 D.D. APSLEY AND M.A. LESCHZINER

Cµ ∼ T0 /t0 ∼ 1/s as s → ∞, where the dimensionless shear parameter s ≡


(k/ε)(2Sij Sij )1/2 ∼ t0 /T0 . In recent years, there has been a gradual movement
away from constant-coefficient (Cµ ≈ 0.09) models to those with strain-dependent
coefficients. This is exemplified by the SST model [23] and in the various non-
linear eddy-viscosity models to be described in the subsections below.

3.2. T HE HYBRID MODELS OF MENTER [23]


Despite its better performance in near-wall flows subject to adverse pressure gradi-
ent than the k–ε model, the k–ω model is observed to be highly dependent on the
free-stream value of ω. Whilst this is not a problem in internal flows, it is clearly
a disadvantage in aerospace applications. In contrast, the k–ε model is apparently
much less sensitive to the free-stream value of ε. A pragmatic approach to these
observations was adopted by Menter [23] who constructed a hybrid k–ω/k–ε model
(the “baseline model”) by adopting a common k–ω formalism, with coefficients in
the ω equation given by
C = F1 C (k−ω) + (1 − F1 )C (k−ε) , (12)
where C (k−ω) are coefficients from the original Wilcox [30] k–ω model and
C (k−ε) are coefficients resulting from the transformation of the standard high-Re
ε equation. F1 is an interpolation parameter such that
F1 = 1 ⇒ k–ω model,
F1 = 0 ⇒ k–ε model (13)
and it is of the form
 1/2 
(k /ω) (v/ω)1/2 [kω/(∇k · ∇ω)]1/2
F1 = F1 , , . (14)
y y y
This last functional form clearly illustrates how the model varies between a k–ω
and k–ε model as the ratio of wall distance y to various turbulence length-scales
changes.
A further advance was the introduction of the shear-stress-transport (SST)
model, for which, in addition to the above hybrid formulation,
 
k a ω
vt = min 1, , (15)
ω F2 
where a is a constant (= 0.31),  is the mean vorticity and
 1/2 
(k /ω) (v/ω)1/2
F2 = F2 , . (16)
y y
Whilst this was originally proposed in order that τ/k should remain constant out-
side the wall layer, it clearly has the same effect as that remarked in Section 3.1,
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 89

namely that Cµ should vary as T0 /t0 or the reciprocal of the strain rate for large
strains. It is not clear why Menter should have used mean vorticity rather mean
strain, or a combination of the two. There is no difference in boundary layers
(∂U/∂y being the dominant component), but a clear distinction arises in largely
irrotational flow regions, such as near impingement points.

3.3. N ON - LINEAR EDDY- VISCOSITY MODELS

Linear EVMs have a single degree of freedom (the scalar vt ) and, therefore, can
only be justified in flows where a single Reynolds-stress component is dynamically
significant; in fully-attached, near-equilibrium shear flows, for example. However,
by generalising the stress-strain relationship, it is possible to mimic the response
of turbulence to complex strain, whilst still retaining a local, one- or two-equation
turbulence model.
The general stress-strain relationship may be written in non-dimensional form
(with second-rank tensors denoted by bold type, T, contracted products indicated
by juxtaposition as for matrix multiplication, (AB)ij = Aik Bkj , and traces denoted
by {T} = Tkk ):

a = f(s, ω). (17)

aij = ui uj /k − (2/3)δij are the components of the anisotropy tensor a, whilst s


and ω are, respectively, the dimensionless irrotational and rotational strain tensors:
 
1 ∂Ui ∂Uj 2 ∂Uk
sij = t0 + − δij ,
2 ∂xj ∂xi 3 ∂xk
 
1 ∂Ui ∂Uj
ωij = t0 − − 2εij k k . (18)
2 ∂xj ∂xi
Here, k represents any system rotation. The main constraints placed upon f
are those of symmetry and zero trace, but other properties associated with the
Reynolds-stress tensor – for example, realisability – may also be imposed.
The Cayleigh–Hamilton theorem dictates that there are only ten tensorially-
independent, symmetric, traceless, second-rank tensor products of s and ω [25], so
that, with one particular basis (the choice of which is discussed below), the general
non-linear stress-strain relationship may be written

a = −c1 s
   
1 2 1 2
+ c2 s − {s }I + c3 (ωs − sω) + c4 ω − {ω }I
2 2
3 3
 
2
+ c5 ω2 s + sω2 − {ω2 }s − {ωsω}I + c6 (ωs2 − s2 ω)
3
90 D.D. APSLEY AND M.A. LESCHZINER
   
1 2 2 2 2
+ c7 ω s + s ω − {ω } s − {s }I − {s ω }I
2 2 2 2 2 2
3 3
 
1
+ c8 s2 ωs − sωs2 − {s2 }(ωs − sω)
2
 
1 2
+ c9 ωsω − ω sω − {ω }(ωs − sω)
2 2
2
+ c10 (ωs2 ω2 − ω2 s2 s), (19)

where successive lines contain tensor bases of increasing degree in s and ω, and
the coefficients ci are functions of the irreducible invariants {s2 }, {ω2 }, {s3 }, {sω2 }
and {s2 ω2 }. In 2-d, incompressible flow, one finds that
1 2
s2 = (s11
2
+ s12
2
)I2 = {s }I2 ,
2
1
ω2 = −ω12
2
I2 = {ω2 }I2 , (20)
2
where I2 = diag(1, 1, 0), so that, with the (slightly complex) basis chosen here, the
cubic and higher-degree terms vanish. Therefore, a quadratic model is sufficient to
compute a 2-d (incompressible) flow. Note also that, in incompressible flow,

P /ε = −{as} (21)

and the quadratic terms do not contribute to the production of turbulence energy.
As pointed out by a reviewer, all ten basis elements are needed for mathemat-
ical completeness in 3-d flows. However, the stress-strain relationship can generate
normal-stress anisotropy and provide a qualitatively correct response of turbulence
to curvature and swirl if terms up to the cubic are retained [7]. If one adheres to the
principle that pure rotation (s = 0, ω 6 = 0) generates no anisotropy, then c4 = 0.
At cubic level, the number of remaining free coefficients (c1 , c2 , c3 , c5 , c6 ) then
matches the five degrees of freedom of the symmetric, traceless tensor aij . This is
the level marked by the horizontal line in Equation (19).
Different terms in (19) can be identified as yielding particular strain sensitivity.
The first term in (19) corresponds to a linear EVM, with c1 = Cµ fµ . The Craft
et al. [7] and Apsley and Leschziner [1] models include parts proportional to {s2 }s
and {ω2 }s (here regarded as parts of the linear term). In curved shear flow,
 
∂U U k 2
{s } − {ω } = −2
2 2
, (22)
∂R R ε
where R is the local radius of curvature. The inclusion of these components, there-
fore, yields an important sensitivity to mean streamline curvature. For simple shear,
the linear term makes no contribution to the normal stresses, but the quadratic terms
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 91

give
1
a11 = (c2 + 6c3 − c4 )σ 2 ,
12
1
a22 = (c2 − 6c3 − c4 )σ 2 ,
12
1
a33 = − (c2 − c4 )σ 2 , (23)
6
where σ = (k/ε)(∂U/∂y), and are responsible for normal-stress anisotropy. In 3-d
flow, the cubic terms can be shown to yield a sensitivity to swirl.
Results for three models are reported here: Gatski and Speziale [11], Craft et
al. [7] and Apsley and Leschziner [1]. The first is quadratic and is a high-Re k–ε
model; the last two are cubic, low-Re k–ε closures. The coefficients in the stress-
strain relationships are given in the Appendix. For details on the ε equations the
reader is referred to the original papers.

3.4. R EYNOLDS - STRESS TRANSPORT MODELS

Differential stress models (DSMs) solve modelled forms of the exact transport
equations:
D ∂
ui uj = dij k + Pij + 8ij − εij . (24)
Dt ∂xk
Whilst the production terms Pij (which transfer energy from the mean flow) and
the advection terms (which incorporate history effects) are exact, the modelling
dilemma is increased over that for the single turbulent scalar k by the need to
model diffusion dij k and dissipation εij for each stress component, as well as the
pressure-strain correlation 8ij , which redistributes energy amongst components.
In addition, numerical stability can be compromised by the lack of a natural dif-
fusivity, although this can be offset by the addition and subtraction of “effective”
eddy viscosities in the momentum equations.
The majority of DSMs use fairly simple models for diffusion, most notably the
generalised gradient-diffusion hypothesis:
 
k ∂
dij k = vδkl + Cs uk ul ui uj , (25)
ε ∂xl
and incorporate any anisotropy of εij into 8ij , so that
2
εij = δij ε. (26)
3
A notable exception is the model of Jakirlić and Hanjalić [16]. The major differ-
ences between DSMs are the modelling of 8ij and the transport equation for (or
other means of specifying) ε.
92 D.D. APSLEY AND M.A. LESCHZINER

Based on the exact integral expression for 8ij , the pressure-strain correlation is
typically decomposed into “slow”, “rapid” and “wall-reflection” parts:

8ij = 8(1) (2) (w)


ij + 8ij + 8ij . (27)

Whereas the exact expression involves volume and surface integrals, the modelled
form makes use of the local stress and strain tensors. Models examined here can be
written in the form
 
0 2 2
8 /ε = −C1 a − C1 a − {a }I ,
(1) 2
3
 .  .
1 1
8 /ε = −C2 P − {P}I
(2)
ε − C3 D − {D}I ε − C4 s,
3 3
 
2
= c01 s + c11 sa + as − {as}I + c12 (ωa − aω),
3
3 3
8(w)
ij /ε = 8̃kl nk nl δij − 8̃ik nj nk − 8̃j k ni nk , (28)
2 2
where
 ε  k 3/2 /ε
8̃(w)
ij = C1(w) ui uj + C2(w) 8(2)
ij f, f = ,
k Cl yn
 
∂Uk ∂Uk
Dij = − ui uk + uj uk . (29)
∂xj ∂xi
With the exception of the Speziale et al. [28] model, all closures tested here use
simple return-to-isotropy (C10 = 0) and isotropisation-of-production (C3 = C4 =
0) approximations, in conjunction with wall-reflection corrections.
Whereas the 8(1) part promotes a return to isotropy, the 8(w)
ij part is designed
to yield the correct anisotropy in the equilibrium layer near walls. The presence
of wall-normal vectors is undesirable, however, and much effort has been made to
eliminate this term, through the use of more complex models for 8(2)ij , and replace
the ni by wall-direction/distance indicators based on the local turbulent length-
scale gradient [5].
Results with four models are reported here: the high-Re models of Gibson and
Launder [12] and Speziale et al. [28] and the low-Re closures of Jakirlić and
Hanjalić [16] and Wilcox [31]. The Speziale et al. closure employs a non-linear
pressure-strain model, with no wall-reflection terms. The Jakirlić and Hanjalić
closure includes anisotropic dissipation and pressure-strain coefficients which are
functions of the stress and dissipation “flatness” factors, A and E, respectively.
The Wilcox multiscale model is ω-based and constructed on the premise that one
can partition the turbulence spectrum into large-scale, energy-bearing eddies and
small-scale, isotropic, dissipative eddies. This construction can be transformed
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 93

into a DSM, supplemented by transport equations for ω and the upper-partition


turbulence energy k U .

4. Results and Analysis


4.1. G RID DEPENDENCE AND INTER - CODE COMPARISON

Important conclusions about turbulence models can only be drawn after all signific-
ant sources of numerical contamination have been eliminated. Thus, an important
aspect of collaborative testing is to ensure grid-independent and code-independent
solutions.
Computations at UMIST were performed with the in-house finite-volume code
STREAM, employing the SIMPLE pressure-correction algorithm with a cell-
centred, co-located storage arrangement and a second-order, three-point, TVD
advection scheme. The time-independent, incompressible-flow equations were
solved iteratively by standard line-iteration/tri-diagonal matrix solvers. Solution
for the pressure was accelerated by ensuring inflow/outflow mass conservation at
each outer iteration and employing block-correction techniques prior to directional
sweeps in the pressure-correction equation. Comparable pressure-correction meth-
ods were used by Loughborough University and Rolls-Royce. ARA and DERA
used their own SAUNA code, which employs cell-vertex storage, Jameson-style
artificial dissipation and explicit time-marching to steady state. BAE SYSTEMS
used a density-based, Jameson-type, time-dependent, compressible code at very
low Mach number.
Figure 3 shows profiles of streamwise mean velocity and Reynolds shear stress
computed on the primary grid (292 × 96 control volumes) and grids where the
mesh density was halved or doubled in both coordinate directions. Computations
were performed with the Wilcox [30] linear k–ω model, which is observed to give a
moderate degree of separation. A fourth mesh, with reduced near-wall cell height,
was also investigated to examine sensitivity to the numerical implementation of the
ω boundary condition. The results shown confirm that, for low-Re models, the ad-
opted grid was sufficient for numerical accuracy. However, for models employing
wall functions to bridge the viscous sublayer, the minimum y + criterion restricts
the level of grid refinement permissible near boundaries.
Inter-code comparisons on the same grid (Figures 4 and 5) show a moderate
degree of scatter. The figures show mean-velocity profiles computed by different
organisations with linear k–ε [19] and linear k–ω [30] models. Included in the
latter are results from the BAE SYSTEMS k–g model, since, for internal flows,
this is a direct transformation of the Wilcox model. The findings are similar to
those of a recent ERCOFTAC workshop [15]; namely, that the level of agree-
ment between different organisations is substantially better with the k–ω model
than with the Launder–Sharma scheme. This probably reflects the complexity of
the length-scale-determining transport equation. Whilst the Wilcox model makes
no concession to special viscous damping terms, the Launder–Sharma closure
94 D.D. APSLEY AND M.A. LESCHZINER

Figure 3. Grid-dependence tests with Wilcox k–ω model: (a) mean velocity; (b) shear stress.

includes “numerically-awkward” second derivatives of velocity. Extrapolated to


many of the more “advanced” turbulence models, this issue of model complexity
becomes significant.

4.2. L INEAR EDDY- VISCOSITY MODELS

Figure 6 shows the development of mean-velocity and Reynolds shear stress com-
puted along the diffuser with representative k–ε [19] and k–ω [30] models, and
with the SST model [23]. Comparisons with other high and low-Re k–ε [20, 22]
and k–ω [32] models show the results for this test case to be insensitive to the
near-wall viscous-damping terms, although not to the high-Re source/production
coefficients Cε1 and Cε2 in the ε equation.
The k–ε model fails to predict separation, producing a nearly symmetric mean-
velocity profile across the diffuser. The k–ω model predicts separation at x/H ≈ 5,
but only weak recirculation. The best overall velocity profiles are recorded with
the SST model, but in this case separation occurs far too early (x/H ≈ 2, com-
pared with an experimental value of 7). The Obi et al. data set was not extended
sufficiently far downstream to examine recovery after reattachment; however, com-
parison with the related Buice and Eaton data shows an insufficient rate of recovery,
a typical finding with Reynolds-averaged models applied to separated flow.
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 95

Figure 4. Comparison of k–ε calculations by different organisations.

Comparison with experiment in Figure 6 shows that all models predict an an-
omalous retardation of the boundary-layer flow and peak in the shear stress (and,
indirectly, the other stress components) following the first corner. This is not seen
in the LES results [10]. A possible suggestion is that the near-wall shear stress fol-
lowing the corner might be influenced by periodic “flapping” or coherent structures
following minor separation at the corner. This organised motion is not captured
by any Reynolds-averaged Navier–Stokes (RANS) model, but is a very effective
mixer. Because the modelled shear stress is too low at the wall at the first corner,
the boundary layer tends to be closer to separation (stronger inflexion and larger
displacement). This leads to higher velocity gradients away from the wall (due to
excessive displacement – an inviscid feature) and hence to the anomalous turbu-
lence peak. As the flow progresses, this excessive shear stress leads to excessive
velocity-profile flattening, by turbulent mixing, and hence to the reduced back-
flow velocity. This behaviour, which is apparent to a greater or lesser extent in all
model results, is rooted in the models’ inability to capture crucial processes just
downstream of the first corner.
A more detailed analysis of the SST calculation is shown in Figure 7. In Fig-
ure 7a the interpolation factor F1 (Equation (14)) is plotted, with a contour marking
the value 0.5. Since F1 = 0 and F1 = 1 correspond to k–ε and k–ω models, re-
96 D.D. APSLEY AND M.A. LESCHZINER

Figure 5. Comparison of k–ω calculations by different organisations.

spectively, it is evident that the model is of k–ω type over a greater part of the flow,
and, in particular, within the important near-wall region. This observation, which is
typical of fully-developed internal flows, but not external flows with a free stream,
explains why calculations with Menter’s baseline model [23] (the hybrid model
without the eddy-viscosity limiter) gives results almost identical to the standard k–
ω model. Figure 7b shows the value of the limiter F2 /aω (Equation (15)), a ratio
of turbulent to shear time scales. The limiter only has an effect when its value is
greater than unity. The figure shows that this occurs in a relatively small region of
the flow: crucially, in the boundary layer just downstream of the first corner, where
it causes separation.
Figure 8 shows derived quantities for eddy-viscosity models: the kinematic
eddy viscosity vt , turbulent kinetic energy k and a “mixing” length l, defined
(independently of the second turbulence variable) by vt = (Cµ1/4 k 1/2 )l. The nor-
malisation is such that l = κyn in the log layer. The plotted models are linear
k–ε, k–ω and SST k–ω models, together with the NLEVM of Craft et al. [7]. Two
features are worthy of note from the inflow profiles (x/H = −11); namely, that
the eddy viscosity is fairly small (the maximum value of vt /v is only about 50)
and that, whilst l/κyn is almost constant across much of the domain, its value is
slightly less than 1.0. The peaks in turbulence energy increase and move towards
the centre of the channel as the flow progresses along the diffuser. Peaks in l/κyn
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 97

Figure 6. Linear eddy-viscosity model calculations of mean velocity and Reynolds shear
stress.

appear relatively close to both walls, being much larger for k–ε than k–ω models.
The length scale predicted by the linear k–ε model increases rapidly along the
diffuser. In contrast, the length-scale ratio for the NLEVM remains limited in a
range roughly 1.0–2.0.

4.3. A NISOTROPY- RESOLVING CLOSURES

Figure 9 shows mean-flow and Reynolds-stress profiles computed with three


NLEVMs: the high-Re quadratic model of Gatski and Speziale [11] and the low-Re
cubic models of Craft et al. [7] and Apsley and Leschziner [1], with the SST model
included for comparison. All models predict separation, with good reproduction
of the maximum velocity along the diffuser, but underprediction of reversed-flow
velocity. As anticipated, a non-linear stress-strain relationship leads to a qualitat-
ively correct anisotropy, with u2 > v 2 . However, whereas the experimental data
reveal the turbulence energy levels to increase monotonically along the diffuser,
the eddy-viscosity closures tend to exaggerate the variations in stresses near the
entrance region and underpredict the overall turbulence level downstream. Many
of the deficiencies noted for linear EVMs in the post-corner region are apparent
here too. An early tendency to a thickening of the boundary layer creates an en-
hanced velocity gradient away from the wall, but the turbulent energy generated
98 D.D. APSLEY AND M.A. LESCHZINER

Figure 7. SST model parameters: (a) interpolation factor F1 ; (b) eddy-viscosity limiter
F2 /aω.

in this region causes greater mixing further downstream. The excessive degree of
anisotropy and anomalous peak in u2 are a direct consequence of the large velocity
gradients away from the wall just downstream of the corner.
Whilst the non-linear k–ε models tested clearly show improved predictive
capabilities relative to their linear counterparts, this cannot be attributed to their
non-linear stress-strain relationship per se, since all feature strain-dependent coef-
ficients. Some effort has been made, therefore, to distinguish the effects of a
strain-dependent Cµ from those due to non-linear terms in the Craft et al. model
(Figure 10). Velocity and stress profiles computed with the complete model and
with the successive removal of quadratic and “curvature” terms (({s2 } − {ω2 })s –
regarded as cubic in the original paper) show that, whereas the quadratic terms
are clearly required to predict anisotropy, the major effect on mean-flow dynamics
is due to the linear term and, in this test case, the curvature-related terms are un-
important. The inclusion of quadratic terms increases anisotropy and, thereby, the
gradients ∂u2 /∂x and ∂v 2 /∂y in non-equilibrium shear flow, leading to a slightly
greater tendency to separation.
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 99

Figure 8. Derived quantities for eddy-viscosity models; (a) eddy viscosity; (b) turbulent
kinetic energy; (c) mixing length.
100 D.D. APSLEY AND M.A. LESCHZINER

Figure 9. Non-linear eddy-viscosity model calculations of mean velocity and Reynolds


stresses (SST model included for comparison).
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 101

Figure 10. Relative effect of variable Cµ and non-linear terms in Craft et al. cubic k–ε model.
102 D.D. APSLEY AND M.A. LESCHZINER

Figure 11. Differential stress model calculations of mean velocity and Reynolds stresses.

Figure 11 shows velocity and stress profiles for various differential stress mod-
els: the high-Re models of Gibson and Launder [12] and Speziale et al. [28],
and the low-Re closures of Jakirlić and Hanjalić [16] and Wilcox [31]. There is
a clear distinction between the wall-function calculations, for which the mean-
flow predictions are comparable to those with anisotropic eddy-viscosity closures,
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 103

and the low-Re calculations, which fail to predict separation. In this instance,
despite its manifest incompatibility with non-equilibrium boundary layers, the
wall-function approach, which fixes the turbulent length-scale and anisotropy com-
ponents in the near-wall region, clearly outperforms low-Re schemes calibrated in
near-equilibrium flows.
The relatively poor performance of the Jakirlić–Hanjalić model is surprising,
for it is, in essence, a low-Re extension of the Gibson–Launder model (albeit with
slightly different coefficients). Moreover, the model originators have themselves
reported favourable predictive properties of the model over a wide range of flows,
including separation. Yet, the present results are supported, qualitatively, by cor-
responding weaknesses observed in shock-affected separation [3] with an entirely
different numerical procedure based on a Riemann solver. There are, however,
variants of the Jakirlić–Hanjalić model, distinguished by additional terms in the
ε equation. In particular, Hanjalić et al. [13] use a variant which includes two extra
source terms, intended to improve the performance of the model in non-equilibrium
flows:

Sε0 = Sl + Sn , (30)

where
 3/2 
εε̃ k
Sl = max[(γ − 1)γ , 0] A,
2 2
γ = ∇ , (31)
k 2.5ε
 
∂Us ∂Un ε
Sn = −1.16 u2s + u2n . (32)
∂xs ∂xn k

Sl is introduced to compensate for excessive growth of the length scale, increasing


dissipation and hence reducing turbulence when the length-scale gradient γ in-
creases faster than its equilibrium-layer value of unity. Sn increases the response of
dissipation to irrotational strains – in particular that associated with adverse pres-
sure gradient. A form of this correction was originally proposed in [14], but strictly
for attached flows. The extra term is not frame-invariant and fails in separated flows
unless it is re-interpreted [21] to apply in a streamline-oriented coordinate system,
with coordinates xs and xn parallel and normal to the local mean velocity vector,
(not easily extensible to three dimensions) and stresses rotated accordingly. The
effect of these terms is considerable (Figure 12). Although Sl is non-negative, it is
comparatively small and has negligible effect on the mean flow. A very significant
effect, however, is produced by the addition of Sn , which is a large positive term in
an adverse pressure gradient, ∂Us /∂xs < 0; (note that, in 2-d incompressible flow,
∂Un /∂xn = −∂Us /∂xs and, usually, u2s > u2n ). The addition of the Sn term leads to
separation and reattachment at x/H ≈ 7 and 26, respectively, in good agreement
with experiment. Moreover, it reduces the anomalous peak in the shear stress just
downstream of the first corner, although the reason for this is unclear.
104 D.D. APSLEY AND M.A. LESCHZINER

Figure 12. Effect of additional ε source terms in the Jakirlić and Hanjalić differential stress
model.

A strong dependence on the ε equation is also found in the high-Re models.


The Gibson–Launder and Speziale et al. models differ only in the models for the
pressure-strain correlation and in the ε-equation coefficient Cε2 :1.92 as against
(an assumed value of) 1.83. Figure 13 shows that it is the latter coefficient which
accounts for almost the entire difference between the two calculations.
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 105

Figure 13. Effect of varying Cε2 in the SSG differential stress model.

In general, the high-Re DSM predictions of the Reynolds stresses are better
than those with NLEVMs. Whilst both predict comparable u2 /v 2 anisotropy, the
DSMs correctly predict the general increase in turbulence energy along the diffuser.
However, there are abnormal levels of normal stress and an anomalous peak in
Reynolds shear stress apparent at the first measurement station.

4.4. P RESSURE AND SKIN - FRICTION COEFFICIENTS

Surface pressure and skin-friction coefficients are plotted for representative linear
EVM, NLEVM and DSM in Figure 14.
The pressure coefficient cP (which is only available for the plane wall from
Obi et al.’s data, but for which computations suggest that the distribution is not
significantly different on the opposite wall) reflects the shape of the recirculating
flow region (i.e. the distance between plane surface and separation streamline). The
plot suggests that the eddy-viscosity closures predict the depth of the recirculating
flow fairly well, but that the SSG model (with Cε2 = 1.83) yields too deep a region
of backflow and too great a pressure loss due to mixing along the diffuser.
Obi et al. provide no data for skin friction. However, such data were obtained
by Buice and Eaton to assist in establishing separation and reattachment points. A
106 D.D. APSLEY AND M.A. LESCHZINER

Figure 14. Pressure and skin-friction coefficients: (a) pressure coefficient on plane surface
(experimental data from Obi et al.); (b) skin friction on inclined surface (experimental data
from Buice and Eaton).
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 107

comparison of experimental and computed values of cf (on the inclined surface)


is shown in Figure 14b, but it must be stressed again that the measured data arise
from a different experiment. The region of backflow is identified (for a Reynolds-
averaged calculation) by negative cf . This parameter is far more sensitive to the
turbulence closure than cP . Calculations with the SST model [23] and the DSM of
Speziale et al. [28] yield far too early separation, whilst the cubic eddy-viscosity
model of Craft et al. [7] gives separation at a location consistent with the experi-
mental data. Interestingly, the last model predicts a tiny region of separation right
at the corner, a feature also revealed by the LES results, although neither the Obi
et al. or Buice and Eaton data sets were able to clarify this.
Pressure coefficients were here referred to a reference datum at x/H = −4.
One feature which has not been explained (but is entirely consistent with other
independent computations [15]) is the high pressure gradient predicted by all mod-
els in the fully-developed flow upstream of the diffuser. Unfortunately, there were
neither experimental measurements of skin friction in the upstream duct from either
the Obi et al. or Buice and Eaton data sets, nor LES data, to investigate this further.

5. Conclusions
This paper conveys a flavour of the difficulties involved in arriving at a reas-
onably secure statement on the relative performance of turbulence models in a
non-trivial, though by no means very complex, flow. In fact, it hides considerably
more than it reveals in terms of the magnitude of the interaction between partners
contributing to this exercise and the numerous iterations needed to illuminate the
sources of inconsistencies and inexplicable differences which characterised early
comparisons. Hence, a first conclusion of the study is that a closely-coordinated
and properly-supported programme, with careful attention to numerical issues and
boundary conditions, is a key to gaining an objective view of the intrinsic predictive
capabilities of alternative closure strategies.
A second conclusion is that no class of models can be said, without qualification
and purely on the basis of agreement with experiment, to be clearly superior or in-
ferior to others. Differential stress models are, of course, fundamentally superior to
eddy-viscosity models in terms of their ability to represent the complex interaction
between different types of strains and all the stress components, but this does not,
in itself, secure superior agreement with flow properties of practical interest in any
but the simplest flows.
In judging the relative merits of models, one issue which poses some uncer-
tainty is the possibility, conjectured on the basis of defects observed consistently
across the entirely range of models investigated, of periodic shear-layer instabilities
provoked by the upstream corner of the diffuser. The existence of such a process
would explain a general failure of all models to resolve the initial development of
the shear layer along the inclined diffuser wall.
108 D.D. APSLEY AND M.A. LESCHZINER

In terms of relative performance, the clearest conclusion is that “standard” lin-


ear eddy-viscosity models, of whatever ilk, which do not contain strain-dependent
coefficients and/or other strain/vorticity-related corrections, fail to represent ad-
equately the present flow, characterised by a strong adverse pressure gradient
that provokes separation. Linear models can and do give satisfactory mean-flow
predictions, provided that they include strain/vorticity-dependent corrections and
are carefully calibrated. The SST k–ω model is an example, although the present
calculations show that some of the predictive advantages are gained at the expense
of defects elsewhere. Thus, while the SST model correctly predicts the general
evolution of the diffuser flow, it does so at the expense of premature separation.
Certainly, no linear model is able to represent potentially important effects arising
from anisotropy and curvature strain.
Non-linear eddy-viscosity models all pursue the same objectives, namely to
return the correct level of anisotropy, to capture the influence of curvature and
to represent realistically the distinctively different response of turbulence to rota-
tional and irrotational straining. However, there are important differences between
models, in terms of their construction and calibration, and it is not at all surprising,
therefore, that they are often found to perform differently in flow conditions that
are remote from those used as a basis of their calibration. In the present flow, it
has been shown that the contribution of the non-linear terms is clearly subordinate
to the strain-sensitised eddy-viscosity coefficient. While the quadratic terms are
instrumental in resolving anisotropy, this anisotropy is not dynamically significant.
It is the inclusion of the strain-dependent eddy-viscosity coefficient that is respons-
ible for all non-linear models returning a broadly correct response of the flow to
the adverse pressure gradient, although none gives the appropriate level of reverse
flow.
Differential stress models have been found to display the widest range of pre-
dictive performance, and some models have given disappointing results. This might
be surprising, at first sight, for this class of models rests most firmly on fundamental
foundations. However, it must be born in mind that the pressure-strain interaction is
a crucially important element influencing the predictive performance. Specifically,
the level of the normal stresses is highly sensitive to this process, and so is the shear
stress, the latter through both normal-stress-induced shear-stress generation and
the shear-stress decay term associated with isotropisation. Hence, the intervention
of the pressure-strain term, the modelling of which is challenging, is an important
obstacle to deriving the full potential of the greater fundamental strength of second-
moment closure. The modelling problems are further compounded by the need
to secure, in low-Re model variants, the correct wall-asymptotic behaviour of all
stress and dissipation components.
Specifically in relation to the present flow, it has to be acknowledged that the
second-moment models investigated herein do not fare uniformly better than sim-
pler closure strategies. Curiously, high-Re variants, coupled with log-law based
wall functions, returned better solutions than low-Re models which account for
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 109

non-equilibrium processes in the semi-viscous sublayer, and this is perhaps a


reflection of the difficulties mentioned earlier in relation to modelling the wall-
asymptotic behaviour. It has also been shown, specifically by reference to the
comparison between the Gibson and Launder [12] and the Speziale et al. [28] mod-
els, that the solutions are highly sensitive to minor changes in the coefficients of
the ε equation, and that this, more than anything else, is responsible for significant
changes in predictive behaviour. Non-equilibrium modifications to the ε-transport
equation in the Jakirlić and Hanjalić [16] model – in particular to distinguish the
effects of rotational and irrotational strains – produced significant improvements,
both in capturing separation and returning correct levels of stress along the entire
length of the diffuser.

Acknowledgements
This paper would not have emerged without the substantial and sustained efforts
made by the partners of the VoTMATA project, in terms of the provision of com-
putational solutions, technical discussions and financial support. Thanks go, in
particular, to Dr. Alan Gould and Dr Stephen Moir (BAE SYSTEMS), Professor
Jim McGuirk and Dr. R.G.M. Hasan (Loughborough University of Technology),
Dr. Nadji Chioukh (Rolls-Royce), Dr. Nick May (ARA) and Dr. Tony Hutton
(DERA).
The financial support of EPSRC, BAE SYSTEMS, Rolls-Royce and DERA is
gratefully acknowledged. Some of the computations included in the paper have
been performed on the SGI Origin 2000 computer (Fermat) at the CSAR national
facility at Manchester Computing, with resources granted by EPSRC.

Appendix: Coefficients in Non-Linear Stress-Strain Relationships


The non-zero coefficients in the general stress-strain relationship (19) are set out
for three models below.

Gatski and Speziale [11] – as regularised by Speziale and Xu [29]:


(1 + 2ζ 2 )(1 + 6η5 ) + 53 η2 α 
1
c1 = ,
(1 + 2ζ )(1 + 2ζ + η + 6β1 η ) 2
2 2 2 6

(1 + 2ζ 2 )(1 + η4 ) + 23 η2  α3,2 
c2,3 = , (33)
(1 + 2ζ 2 )(1 + 2ζ 2 + β3,2 η6 ) 2

1
η = (α3 /α1 )(sij sij )1/2 , ζ = (α2 /α1 )(ωij ωij )1/2,
2
α1 = 0.227, α2 = 0.0424, α3 = 0.0397,
β1 = 7.0, β2 = 6.3, β3 = 4.0. (34)
110 D.D. APSLEY AND M.A. LESCHZINER

Craft, Launder and Suga [7]:


 
c, 1 = Cµ fµ 1 + 10Cµ2 s 2 − ω2
,(c2 , c3 , c4 ) = (−0.4, 0.4, −1.04)Cµ fµ
c, 6 = 80Cµ3 fµ (35)

0.3[1 − exp(−0.36e0.75s̃ )]
Cµ = ,
1 + 0.35s̃ 3/2
fµ = 1 − exp[−(Rt /90)1/2 − (Rt /400)2 ],
s = (2sij sij )1/2 , ω = (2ωij ωij )1/2 , s̃ = max(s, ω). (36)

Apsley and Leschziner [1]:


"  #

(−a12 ) 1 + 13 β 2 s 2 − γ 2ω2 /s̃ 2
c1 = ,
s̃ 1 + 13 β 2 − γ 2
∗ ∗ ∗ ∗
6(a11 + a22 ) a11 − a22
c2 = , c3 = ,
s̃ 2 s̃ 2

(−a12 )
(c5 , c6 ) = (−6γ 2 , −6βγ ) 3 , (37)
s̃ 1 + 3 β − γ 2
1 2

s = (2sij sij )1/2 , ω = (2ωij ωij )1/2 ,


1 2f0
= p ,
s̃ σ ∗ + σ ∗2 + 2f0 (f0 − 1)(s 2 + ω2 )
.f0 = 1 + 1.25 max(0.09σ ∗2 , 1), β = 0.222, γ = 0.623 (38)
∗ ∗ ∗
a12 , a11 , a22 and σ ∗ are functions of y ∗ , based on DNS data for fully-developed
∗ ∗ ∗
channel flow. In the high-Re region (i.e. y ∗ → ∞), (a12 , a11 , a22 , σ ∗) =
(−0.3, 1.0, 0.4, 3.33).

References
1. Apsley, D.D. and Leschziner, M.A., A new low-Reynolds-number non-linear two-equation
turbulence model for complex flows. Internat. J. Heat Fluid Flow 19 (1998) 209–222.
2. Baldwin, B.W. and Barth, T.A., One-equation turbulence transport model for high Reynolds
number wall bounded flows. AIAA Paper 91-0610 (1991).
3. Batten, P., Craft, T.J., Leschziner, M.A. and Loyau, H., Reynolds stress transport modelling for
compressible aerodynamic flows. AIAA J. 37 (1999) 785–796.
4. Buice, C.U. and Eaton, J.K., Experimental investigation of flow through an asymmetric plane
diffuser. Report TSD-107, Department of Mechanical Engineering, Stanford University (1997).
ADVANCED TURBULENCE MODELLING OF SEPARATED FLOW IN A DIFFUSER 111

5. Craft, T.J., Developments in a low-Reynolds-number second-moment closure and its applica-


tion to separating and reattaching flows. Internat. J. Heat Fluid Flow 19 (1998) 541–548.
6. Craft, T.J. (ed.), Proceedings of the 7th ERCOFTAC/IAHR/COST Workshop on Refined
Turbulence Modelling, 28–29 May. UMIST, Manchester (1998).
7. Craft, T.J., Launder, B.E. and Suga, K., Development and application of a cubic eddy-viscosity
model of turbulence. Internat. J. Heat Fluid Flow 17 (1996) 108–115.
8. Craft, T.J., Launder, B.E. and Suga, K., Prediction of turbulent transitional phenomena with a
nonlinear eddy-viscosity model. Internat. J. Heat Fluid Flow 18 (1997) 15–28.
9. Durbin, P.A., Separated flow computations with the k–ε–v 2 model. AIAA J. 33 (1995) 659–664.
10. Fatica, M., Kaltenbach, H.-J. and Mittal, R., Validation of large-eddy simulation in a plane
asymmetric diffuser. In: Annual Research Briefs. Center for Turbulence Research, Stanford
(1997) pp. 23–36.
11. Gatski, T.B. and Speziale, C.G., On explicit algebraic stress models for complex turbulent
flows, J. Fluid Mech. 254 (1993) 59–78.
12. Gibson, M.M. and Launder, B.E., Ground effects on pressure fluctuations in the atmospheric
boundary layer. J. Fluid Mech. 86 (1978) 491–511.
13. Hanjalić, K., Jakirlić, S. and Hadžić, I., Expanding the limits of “equilibrium” second-moment
turbulence closures. Fluid Dynam. Res. 20 (1997) 25–41.
14. Hanjalic, K. and Launder, B.E., Sensitising the dissipation equation to irrotational strains. J.
Fluids Engrg. 102 (1980) 34–40.
15. Hellsten, A. and Rautaheimo, P., (ed.), Proceedings of the 8th ERCOFTAC/IAHR/COST
Workshop on Refined Turbulence Modelling, 17–18 June. Helsinki University of Technology
(1999).
16. Jakirlić, S. and Hanjalić, K., A second-moment closure for non-equilibrium and separating high
and low Re number flows. In: Durst, F., Launder, B.E., Schmidt, F. and Whitelaw, J.H. (eds),
Proceedings 10th Symposium on Turbulent Shear Flows. Pennsylvania State University (1995)
pp. 23.25–23.30.
17. Kalitzin, G., Gould, A.R.B. and Benton, J.J., Application of two-equation turbulence models in
aircraft design. AIAA 96-0327 (1996).
18. Launder, B.E., Reece, G.J. and Rodi, W., Progress in the development of a Reynolds-stress
turbulence closure. J. Fluid Mech. 68 (1975) 537–566.
19. Launder, B.E. and Sharma, B.I., Application of the energy-dissipation model of turbulence to
the calculation of flow near a spinning disc. Lett. Heat Mass Transfer 1 (1974) 131–138.
20. Launder, B.E. and Spalding, D.B., The numerical computation of turbulent flows. Comput
Meth. Appl. Mech. Engrg. 3 (1974) 269–28.
21. Leschziner, M.A. and Rodi, W., Calculation of annular and twin parallel jets using various
discretisation schemes and turbulence-model variations. J. Fluids Engrg. 103 (1981) 352–360.
22. Lien, F.S. and Leschziner, M.A., A pressure-velocity solution strategy for compressible
flow and its application to shock/boundary-layer interaction using second-moment turbulence
closure, J. Fluids Engrg. 115 (1993) 717–725.
23. Menter, F.R., Two-equation eddy-viscosity turbulence models for engineering applications.
AIAA J. 32 (1994) 1598–1605.
24. Obi, S., Aoki, K. and Masuda, S., Experimental and Computational study of turbulent separ-
ating flow in an asymmetric plane diffuser. In: Durst, F., Kasagi, N., Launder, B.E., Schmidt,
F.W., Suzuki, K. and Whitelaw, J.H. (eds), Proceedings 9th Symposium on Turbulent Shear
Flows, Kyoto, Japan, August 16–18. (1993) Paper P305-1.
25. Pope, S.B., 1975, A more general effective-viscosity hypothesis. J. Fluid Mech. 72 (1975)
331–340.
26. Reynolds, W.C. and Kassinos, S.C., One point modelling of rapidly deformed homogeneous
turbulence, Proceedings Osborne Reynolds Centenary Symposium, Manchester. Proc. Roy.
Soc. London, A 451 (1994) 87–104.
112 D.D. APSLEY AND M.A. LESCHZINER

27. Spalart, P.R. and Almaras, S.R., A one-equation turbulence model for aerodynamic flows,
AIAA Paper 92-0439 (1992).
28. Speziale, C.G., Sarkar, S. and Gatski, T.B., Modelling the pressure-strain correlation of
turbulence: An invariant dynamical systems approach, J. Fluid Mech. 227 (1991) 245–272.
29. Speziale, C.G. and Xu, X-H., Towards the development of second-order closure models for
nonequilibrium turbulent flows. Internat. J. Heat Fluid Flow 17 (1996) 238–244.
30. Wilcox, D.C., Reassessment of the scale-determining equation for advanced turbulence models.
AIAA J. 26 (1988) 1299–1310.
31. Wilcox, D.C., Multiscale model for turbulent flows. AIAA J. 26 (1988) 1311–1320.
32. Wilcox, D.C., Simulation of transition with a two-equation turbulence model. AIAA J. 32 (1994)
247–255.

You might also like