You are on page 1of 13

International Journal of Heat and Fluid Flow 42 (2013) 151163

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Flow and performance characteristics of an Allison 250 gas turbine


S-shaped diffuser: Effects of geometry variations
Gavin G. Lee, William D.E. Allan, Kiari Goni Boulama
Department of Mechanical and Aerospace Engineering, Royal Military College of Canada, Kingston, Ontario, Canada K7K7B4

a r t i c l e

i n f o

Article history:
Received 27 March 2012
Received in revised form 31 January 2013
Accepted 12 February 2013
Available online 7 March 2013
Keywords:
S-shaped diffuser
Aircraft inlet duct

a b s t r a c t
The S-shaped diffuser which connects the exit of the compressor to the inlet of the combustion chamber
of the Allison 250 gas turbine has been investigated using the Shear-Stress Transport turbulence model
(SST) and the commercial code ANSYS-CFX. The diffuser geometry includes an initial conical diffuser
which smoothly transitions into a constant cross-section S-duct. The numerical model and setup were
validated using both in-house processed experimental data and experimental data from the literature
on a similar geometry. The stream-wise velocity prole was observed to atten in the initial divergent
section, and then the region of the ow with the highest velocity is pushed toward the outer surface
of the rst bend, with a secondary-ow in the plane of the cross-section. This distortion of the streamwise velocity intensied when the inlet turbulence intensity was decreased or when the Reynolds number was increased. An increase of the Reynolds number also translated into higher static pressure recovery potential and lower wall friction coefcients. Six variations of the diffuser geometry were considered,
all having the same total cross-sectional area ratio and centreline offset. The qualitative results were the
same as those of the Allison 250 diffuser, but unlike the base geometry, all the considered variants
showed separated-ow regions (and reversed-ow regions in some cases) of different sizes and at different locations. The performance indicators for the Allison 250 S-shaped diffuser were the highest overall.
Most interestingly, the current duct geometry outperformed its variant with a cross-sectional area expansion extending over its entire length, which is the most common inlet duct conguration.
2013 Elsevier Inc. All rights reserved.

1. Introduction
Turbulent ows in passages with gradually varying cross-sectional area and bends are present in a large variety of applications and have long been the subject of interest from the
scientic community. 2D and 3D planar diffusers and curved
rectangular ducts have thus been investigated using both the
experimental and numerical approaches, evidencing occurrences
of separated-ow regions, the extents of which depend on the
passage area ratio and the severity of the duct curvature when
applicable, and the Reynolds number (Gullman-Strand et al.,
2004; Cherry et al., 2008; Schneider et al., 2010; Jakirlic et al.,
2010; Gopaliya et al., 2011). As far as axisymmetric geometries
are concerned, Azad (1996) authored a review of three-decades
of experimental research on an 8 conical diffuser at the University of Manitoba. According to the author, diffusers can be
viewed as devices that convert kinetic energy into potential energy, and the efciency of such conversion is the highest for
divergence angles between 6 and 8 for a conical diffuser, a
divergence angle of 11 for a rectangular 2D diffuser, and a
Corresponding author.
E-mail address: Goni.Boulama@rmc.ca (K. Goni Boulama).
0142-727X/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatuidow.2013.02.004

divergence angle of 6 for a square cross-section diffuser. This


efciency is on the other hand dependent upon the nature of
the boundary layer at the inlet of the diffuser. Among other conclusions, the existence of an internal layer was conjectured, and
later veried numerically (Wu et al., 2006). Sparrow et al. (2009)
also conducted research on conical diffusers and noted an inconsistency in the literature, with some sources stating that the
ow does not separate for divergence angles less than 7, while
other sources state that divergence angles greater than 15 cause
ow separation. Furthermore, the authors noted that the Reynolds number effect on the ow separation vs. divergence angle
correlation is often overlooked. Based on these observations,
they numerically investigated a wide range of divergence angles
and Reynolds numbers, and observed ow separation at divergence angles as small as 5 in the laminar regime, while simulations with 1030 divergence angles all predicted ow
separations for Reynolds numbers varying between 500 and
33,000, the extent of the separated-ow region being the greatest at lower Reynolds numbers. The latter conclusion contrasts
with that of Herbst et al. (2007), who observed that the separated-ow region in planar diffusers becomes larger with
increasing Reynolds number, while Yang and Hou (1999) concluded that the Reynolds number effect was ambiguous.

152

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

Nomenclature
a, A, B, C
Cf
CL
Cp
D
KE
NUI
Ps, PT
Re
r, z, x
R, Z, X

see grid nomenclature in Fig. 3 and Table 1


friction coefcient 2sw =qu20
coefcient of total pressure loss 2PT;0  P T =qu20
pressure coefcient 2P s  Ps;0 =qu20
local duct diameter, m
turbulence kinetic energy,Pmp2
/s2
P p
non-uniformity index
v 2 w2 =
u2
2
static pressure, total pressure, N/m
Reynolds number (=qu0D0/l)
space coordinates, m, see Figs. 1 and 3
non-dimensional space coordinates. Radial coordinates
r and z are non-dimensionalized by the local duct diameter, while the axial coordinate x is non-dimensionalized by the total diffuser length

Most important early advances in the study of ows in curved


pipes are due to WR Dean and his discovery that the primary motion along the line of the conduit was accompanied by a secondary
motion in the plane of the cross-section, an outward shift of the region where the primary ow velocity is the greatest, and a decreased mass ow-rate for a given pressure gradient (Dean and
Hurst, 1959). Following suit, several other researchers investigated
the ows in curved ducts and ducts with a series of bends (Talbot
and Wong, 1982; Patankar et al., 1975; Towne, 1984; Rudolf and
Desova, 2007). The study by Talbot and Wong (1982)) specically
examined the nature of the secondary-ow boundary-layer collision, corresponding to the point where the wall shear stress vanishes without changing sign. Patankar et al. (1975) and Towne
(1984) are examples of early applications of computational approaches for complex turbulent ow investigations.
The interest in the 1980s for turbo-prop powered airplanes
came along with the requirement of specic inlet congurations
for the double purpose of ensuring the most uniformly distributed
ow possible and the highest possible pressure recovery at the inlet of the engines (McDill, 1989). Furthermore, Anderson (1991)
noted that the inlet geometry should minimize vortex, wake and
boundary layer ingestion, and minimize ow eld interference
from weapon carriage and ring, landing gear deployment, etc.
According to the author, engine face spatial and temporal ow distortion was one of the least understood problems that early versions of the B70, F-111, F-14, MIG-25, Tornado, Airbus A300, and
other aircraft encountered. The shape optimization study by McDill
(1989) started with an arbitrary superellipse at the diffuser throat
and resulted in a circular cross-section at the compressor face, following a specic area expansion which is, for example, the shape
adopted for the F/A-18 inlet duct (Anderson, 1991). In another
shape optimization study, Saha et al. (2007) compared elliptic,
semi-circular, oval, rectangular and square duct inlet geometries,
while maintaining the same outlet diameter, duct turning angle,
area ratio and Reynolds number. They concluded that the duct
with the semi-circular inlet has the highest pressure recovery
and the lowest pressure loss coefcient, while the elliptic inlet duct
has the least total pressure distortion at the exit plane. The squared
inlet duct performed the poorest overall. Anand et al. (2003) experimentally investigated the effect of the turning angle on the ow in
an S-shaped diffuser with an area ratio of 1.9, at a Reynolds number of 100,000. They observed that the velocity distribution was
more uniform and the pressure recovery was lower for the higher
of the two turning angles tested (i.e. the smaller radius of curvature, or greater centreline offset). A numerical study on an
S-shaped diffuser with a rectangular inlet and a semi-circular outlet reproduced the same trends for the pressure recovery, but not

Tu
u, v, w
U
us
Y+
yp

sw
l
q
m

turbulence intensity (=(1.5KE)0.5/u0)


velocity vector components, m/s
stream-wise velocity non-dimensionalized by the local
mean velocity
p
friction velocity sw =q, m/s
non-dimensionalized wall distance (=ypus/m)
initial wall spacing, m
wall shear-stress (=l(@u/@r)|r=0.5D), N/m2
dynamic viscosity, N s/m2
uid density, kg/m3
kinematic viscosity (=l/q), m2/s

for the ow uniformity (Gopaliya et al., 2007). The latter study also
predicted that an increase in the Reynolds number does not have a
signicant effect on the pressure recovery. Abdellatif et al. (2008)
adopted the numerical approach in their investigation of the effect
of the area ratio on the performance of S-shaped diffusers. Considering an S-shaped diffuser of a constant turning angle and centreline
length, they changed the area ratio from 1 to 1.51, which resulted
in a static pressure recovery variation from 0.19 to 0.40. A further
increase of the area ratio to 1.9 did not signicantly improve the
pressure recovery, but the ow distortion increased. The authors
therefore concluded that for the given constraints, the area ratio of
1.51 constitutes an optimum. Whitelaw and Yu (1993) experimentally studied an S-shaped diffuser tube: the Royal Aerospace Establishment (RAE) series 2129, which has an area ratio of 1.4 and a nondimensionalized offset of 0.3. They used a Reynolds number of
400,000 and two different thicknesses of the boundary layer at the
inlet of the diffuser. In both cases, a region of separation was observed at the inside surface of the rst bend, just upstream of the
junction between the rst and second bends, and it was larger for
the thinner inlet boundary layer. A pair of counter-rotating vortices
was also observed at the exit plane, which increased the non-uniformity of the stream-wise velocity. No signicant pressure recovery
changes were observed for the two inlet boundary layer thicknesses
considered in the study. The RAE2129 was later reexamined by Lee
and Yu (1994) using a Reynolds Averaged NavierStokes (RANS)
method. Despite the simplicity of the method, the calculated mean
ow features reproduced the experimental data in Whitelaw and
Yu (1993) with sufcient accuracy.
This literature survey conrms that the research on turbulent
ows in curved and/or diffusing ducts is topical, and several of
its aspects have been studied using both the experimental and
numerical approaches. RANS models have been used in the literature with a wide variety of prediction accuracy and implementation cost (Gullman-Strand et al., 2004; Schneider et al., 2010;
Gopaliya et al., 2007, 2011; Sparrow et al., 2009; Yang and Hou,
1999; Patankar et al., 1975; Rudolf and Desova, 2007; Saha et al.,
2007; Lee and Yu, 1994; El-Behery and Hamed, 2011). Herbst
et al. (2007) mentioned that a suitable RANS model could result
in a reasonable agreement with experimental data, but cautioned
that the prediction capability is highly model-dependent. Their
own Large Eddy Simulation (LES) study predicted a slightly smaller
separated-ow region than the one measured, as well as delayed
separation and reattachment. Abe and Ohtsuka (2010) also used
the LES approach and observed a signicantly under-predicted
wall friction, resulting in an overly early separation. Much
improved prediction capability was obtained when a hybrid
LES/RANS method was used. On the other hand, Ohlsson et al.

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

(2010) adopted the Direct Numerical Simulation (DNS) approach


and demonstrated its capability to provide accurate and detailed
information about the physics of the ow in a 3D straight diffuser.
However, the cost of the DNS approach is currently prohibitive for
high Reynolds number ows in complex geometries including
cross-sectional area variations and curvatures.
In collaboration with a manufacturer partner, our lab is conducting a comprehensive study of the Allison 250 gas turbine model
(A250). The present study focuses on the turbulent ow in the duct
connecting the exit of the compressor to the inlet of the combustion
chamber of the A250. This duct has an initial conical diffuser shape
that smoothly transitions into a constant cross-sectional area, Sshaped duct. With the exception of preliminary results recently presented at a conference by the present authors (Lee et al., 2012), no
previous work has been reported on this particular geometry in
the literature, most inlet ducts having a continuously expanding
cross-sectional area (McDill, 1989; Anderson, 1991; Saha et al.,
2007; Anand et al., 2003; Gopaliya et al., 2007; Abdellatif et al.,
2008; Whitelaw and Yu, 1993; Lee and Yu, 1994). The RANS ShearStress Transport (SST) turbulence model is adopted as the investigation tool. An experimental rig is constructed and used to collect data
that are used for validation. Further validation is performed against
experimental data on the RAE2129 from the literature (Whitelaw
and Yu, 1993). After the main ow features of the A250 S-shaped diffuser duct are calculated, the discussion is continued with a sensitivity analysis on the inlet turbulence intensity and Reynolds number,
and the ow in the A250 S-Diffuser is compared to that in a number
of slightly different geometrical congurations.

2. Experimental setup
The geometry of the Allison 250 model diffuser duct (hereafter
referred to as A250 S-Diffuser) is shown in Fig. 1a. It has a total
length of 426 mm, which includes a 155.2 mm conical section followed by a 270.8 mm constant diameter S-shaped duct. The inlet
and outlet diameters are 46.83 mm and 65.11 mm, respectively,
giving an area ratio of 1.933. The offset between the inlet and
the outlet is 48.2 mm. The top edge of the duct in the view in
Fig. 1a is adjoining the combustion chamber and turbine, and will
herein be referred to as the inner surface of the rst bend, or the
wall at R = 0.5. Accordingly, the bottom edge will be referred to
as the outer surface of the rst bend, or the wall at R = 0.5.
A service-worn sample of the A250 S-Diffuser was obtained as a
test article. A exible traverse system with a 0.8 mm pitch,
equipped with a Pitot-static tube is designed to be placed at three
stream-wise locations (1: entrance; 2: end of conical section; 4:
exit plane of diffuser). Furthermore, pressure measurements could
be taken in the four directions indicated in Fig. 1b. An Omega
PX139 pressure transducer rated for 0.3 psi with an output voltage between 0.25 and 4.25 V is used, in conjunction with a National Instruments Analog/Digital converter (NI cRIO9201). At
each position, measurements are taken over a period of 10 s at

153

100 Hz, which allows for the calculation of a mean value along
with its standard deviation. A Westinghouse 1.5 hp, 3450 rpm motor, mated to a Canadian Blower & Forge Co. centrifugal blower, is
used to generate the desired ow. A 2 m long steel, drawn-overmandrel, tube of inside diameter equal to that of the inlet of the
A250 S-Diffuser is congured to the blower of outlet diameter
4.5 in via a plain reducer and a contraction cone. Honeycomb ow
straighteners are placed at the entrance of this initial tube in order
to minimize instabilities before the ow enters the test-section. In
addition, a 1 m steel extension tube, with an inside diameter equal
to that of the exit of the A250 S-Diffuser, is placed at the exit of the
diffuser in order to prevent outlet conditions from affecting the
ow in the region of interest. This practice is common in both
experimental and numerical approaches (Gullman-Strand et al.,
2004; Cherry et al., 2008; Schneider et al., 2010; Jakirlic et al.,
2010; Sparrow et al., 2009; Saha et al., 2007; Anand et al., 2003;
Gopaliya et al., 2007; Abdellatif et al., 2008; Whitelaw and Yu,
1993; Lee and Yu, 1994; El-Behery and Hamed, 2011; Ohlsson
et al., 2010). Under the considered conditions, a mass ow-rate
of 0.055 kg/s and a Reynolds number of 80,000 are obtained. The
area averaged turbulence intensity at the inlet plane of the component (Station 1) is estimated to be 1.6%. Note that the experimental
rig is, in the long-term, meant to be used in order to study the
propagation of ow instabilities from the compressor to the combustion chamber.
In the following, the radial coordinates z (AA direction) and r
(CC direction) are non-dimensionalized by the local duct diameter,
the axial coordinate x is non-dimensionalized by the total diffuser
length, and the velocity components u, v and w are non-dimensionalized by the local bulk velocity.
In this study, the sensitivity of the ow features and performance
characteristics of the A250 S-Diffuser are discussed by considering
several geometrical variations, see Fig. 2. All geometries in Fig. 2
have the same inlet and outlet diameters and centreline offset as
the A250 S-Diffuser. For the S-Diffuser 1/3 (Fig. 2a), the cross-sectional area expansion occurs between Sections 1 and 3. For the S-Diffuser 1/4 (Fig. 2b), the area expansion extends over the entire duct
length, as it is for the RAE2129 and most inlet ducts. For the S-Diffuser 2/3 (Fig. 2c), the area expansion occurs between Sections 2
and 3. The same nomenclature convention is adopted for the S-Diffuser 2/4 (Fig. 2d), S-Diffuser 3/4 (Fig. 2e) and S-Diffuser 1/23/4 (not
shown). Consistently, the A250 S-Diffuser (Fig. 1a) could be referred
to as S-Diffuser 1/2. Finally, an S-shaped duct (not shown) with a
constant cross-sectional area, a diameter equal to that of the inlet
of the A250 S-Diffuser, and the same centreline offset as the A250
S-Diffuser will also be considered for comparison purposes.

3. Numerical approach
The commercial CFD software ANSYS-CFX is adopted for this
study to solve the steady-state Cartesian coordinates Navier
Stokes equations using a fully conservative, control-volume-based,

Fig. 1. (a) Geometry of the A250 S-Diffuser and (b) pressure traverse paths.

154

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

Fig. 2. (a) S-Diffuser 1/3, (b) S-Diffuser 1/4, (c) S-Diffuser 2/3, (d) S-Diffuser 2/4 and (e) S-Diffuser 3/4.

nite-element method. To ensure robustness, the discretization is


performed with a combination of the upwind and central differencing schemes. The resulting hybrid scheme is rst- and second-order accurate, switching from the latter to the former in
regions of steep gradients, based on the boundedness principle of
Barth and Jespersen (1989). The resolution scheme is second-order
accurate. Pressurevelocity coupling is enforced with a non-staggered grid and the fourth-order-accurate algorithm of Rhie and
Chow (1983).
The Shear-Stress Transport model (SST) is based on the two following transport equations:



@qU j k f
@
@k
P k  b qkx
l rk lt
@xj
@xj
@xj



@U j x
@
@x
aqS2  bqx2
l rx lt
@xj
@xj
@xj
1 @k @ x
21  F 1 qrx2
x @xj @xj

where F1 is equal to zero away from the surface (je model) and
equal to one inside the boundary layer (jx model). The model also
includes a slight amendment to the denition of the eddy viscosity
for a better prediction of the turbulent shear stress. The calculation
of the distance from the wall is achieved by the solution of a Poisson
equation. More details on the SST model could be found in Sparrow
et al. (2009), and Menter (1994), and are not reproduced here for
conciseness. The suitability of the SST turbulence model is discussed in the next section. For each simulation, the convergence

155

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

is judged based on the variations of the scaled residuals of the governing equations (values in the order of or below 106 are desired)
and veried using an integral mass balance.
The wall boundary conditions are the standard no-slip and
impermeability conditions. At the inlet of the computational domain, a uniform velocity prole corresponding to the desired Reynolds number is prescribed. This results in fully-developed ow
conditions at the inlet of the component under study. At the other
end of the computational domain, stream-wise second derivatives
of velocities are assigned a zero value to represent fully-developed
ow conditions. Finally, all simulations are performed with standard air at the constant temperature of 25 C.

Table 1
Characteristics of some of the tested grid resolutions.
Grid resolution

Wall spacing (Y+)

Total

G1
G2
G3
G4
G5
G6

45.4
6.98
1.12
1.12
1.12
0.68

10
25
20
20
20
30

15
55
70
70
70
100

100
100
100
50
150
100

180
180
180
90
270
180

288,108
2,770,893
2,692,188
1,789,148
3,905,648
5,907,168

0.006

3.1. Grid sensitivity

G1
G2
G3

0.004

G4

Cf

G5
G6
0.002

0.25

0.5

0.75

0.75

X*

(a) R* = 0.5
0.006
G1
G2
G3
0.004

G4
G5
G6

Cf

The A250 S-Diffuser used in the physical experiment has been


digitized and imported into ICEM, a grid generator compatible with
the ANSYS-CFX CFD solver. Consistently with the experimental setup, a 1.5 m entrance length (upstream the diffuser) and a 1 m long
extension (downstream the diffuser) are used in the numerical
model. An illustration of the grid topology is shown in Fig. 3. The
grid is uniformly distributed along the component (not shown);
while in the plane of the cross-section the adopted O-grid mesh
is relatively coarse in the central region with a geometric renement as the wall region is approached. This meshing technique allows for an even wall size distribution, and is superior to a purely
hexagonal mesh (Lee and Yu, 1994).
Different grids have been tested, with different stream-wise and
cross-ow resolutions. The most important features of some of
these grids are summarized in Table 1. A represents the number
of nodes along the side of the square central box of Fig. 3, a represents the number of nodes along the line from the corner of the
central square box to the wall, while B and C represent the
stream-wise grid distribution in the conical and S-shaped sections
of the duct, respectively. The calculated friction coefcients for
grids G1G6 are shown in Fig. 4. Rening the cross-sectional plane
resolution from G1 to G2, while maintaining the longitudinal resolution unchanged, results in a very signicant change of the friction
coefcient values (as much as 16% in average at the outer surface
of the rst bend). A further renement of the cross-sectional plane
resolution particularly in the wall region (G2G3) results in an
average variation of the friction coefcient values of 8.4% at the
R = 0.5 wall, and this variation falls to 0.90% only when G3 is
compared to the nest grid G6. On the other hand, the variations
in the friction coefcient values are small (about 0.20% in average

0.002

0.25

0.5

X*

(b) R* = -0.5
Fig. 4. Effects of the grid resolution on the calculated the friction coefcients.

at R = 0.5) when the longitudinal resolution is rened from G4 to


G3 and from G3 to G5, while maintaining the cross-sectional plane
resolution unchanged. In view of these results, and in order to preserve both accuracy and cost, G3 is considered as sufciently wellresolved to be adopted for all subsequent simulations. Notwithstanding, a wall resolution Y+  1.0 was recommended in a similar
diffuser ow with separation in the literature (Rudolf and Desova,
2007; El-Behery and Hamed, 2011; Abe and Ohtsuka, 2010).
3.2. Validation

Fig. 3. Cross-sectional plane grid topology.

The calculated velocity proles using the grid resolution G3, a


Reynolds number of 80,000, a turbulence intensity of 5% at the inlet of the computational domain, and the ANSYS-CFX SST turbulence model are compared with the experimental data in Fig. 5. A

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

0.4

0.4

0.2

0.2

Z*

Z*

156

-0.2

SST

-0.2

Station 1
Station 2
Station 4

-0.4

-0.4
0

0.5

1.5

0.5

U*

(a) AA line

0.4

0.4

0.2

0.2

R*

R*

(a) AA line

0
Station 1
Station 2
Station 4

-0.2

1.5

U*

SST

-0.2

-0.4

-0.4
0

0.5

1.5

0.5

U*

U*

(b) CC line

(b) CC line

1.5

Fig. 5. Development of the stream-wise velocity prole in the A250 S-Diffuser.

Fig. 6. Comparison of the prediction capability by different turbulence models.

very good qualitative and quantitative agreement is observed at all


three measurement stations (see Fig. 1a), and along the two crossow directions AA and CC (see denition in Fig. 1b).
In order to further assess the suitability of the adopted turbulence model, the results shown in Fig. 5 are re-calculated using
the two turbulence models it builds upon: the standard je and
jx models. As discussed in the introduction, several variants of
these models have been used in the literature with different levels
of success (Gullman-Strand et al., 2004; Schneider et al., 2010;
Gopaliya et al., 2007, 2011; Sparrow et al., 2009; Yang and Hou,
1999; Patankar et al., 1975; Rudolf and Desova, 2007; Saha et al.,
2007; Lee and Yu, 1994; El-Behery and Hamed, 2011; Menter,
1994). Fig. 6a shows that the stream-wise velocity distribution
along the AA direction at the exit plane of the A250 S-Diffuser
(see Station 4 results in Fig. 5) is not well predicted by either the
je and jx models, with the latter model qualitatively predicting
the existence of the two overshoot regions next to the wall, and the
former model completely failing to do so. Fig. 6b also shows that
both the je and jx models fail to predict the magnitude of
the velocity prole distortion. Of particular importance is the
velocity gradient predicted by the je model at R = 0.5 on the
one hand, and the one predicted by the SST model at about
R = 0.3 on the other hand. According to El-Behery and Hamed
(2011), the better performance of the SST and jx models compared to the je model (which also requires more computational

time) may be explained by the fact that the former two models account for some near-wall turbulence anisotropy, which is not the
case for the je model. Menter (1994) also noted the very little
sensitivity of the je model to the pressure gradient, and attributed the better performance of the SST model compared to the
jx model to the presence of an additional cross-diffusion term
in the dissipation rate transport equation (last term on the right
hand side of Eq. (2)).
In the absence of published data on the A250 S-Diffuser, the
numerical code has nally also been tested against available experimental data on the RAE2129 diffuser from the literature (Whitelaw and Yu, 1993). The RAE2129 S-shaped inlet duct differs from
the A250 S-Diffuser in that it has a smaller area ratio (1.4), a larger
offset (0.3), and its cross-sectional area expands over its entire
length. The test is conducted for the fully-developed ow inlet condition of Whitelaw and Yu (1993) at a Reynolds number of
400,000. Fig. 7 shows that the development of the stream-wise
velocity prole is well predicted by the model (present calculated
results are shown in the form of velocity vectors), while Fig. 8 also
shows a reasonably good agreement between the experimental
and numerical data; this agreement is by all means at least as good
as the one obtained by the same authors when they adopted the
numerical approach (Lee and Yu, 1994).
In view of the grid resolution sensitivity analysis (Fig. 4), the
experimental validation cases (Figs. 5, 7 and 8) and the turbulence

157

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

Fig. 7. Development of the stream-wise velocity prole in the RAE2129.

0.4

0.2

Cp

(a) AA line
0

-0.2

Present Study Exp


R* = 0.5
R* = 0.0
R* = -0.5

-0.4
-0.1

0.1

0.2

(b) CC line
0.3

X
Fig. 8. Development of the pressure coefcient in the RAE2129 (Experimental data
are from Whitelaw and Yu, 1993).

model discussion (Fig. 6), the numerical model (the SST turbulence
model as imbedded in ANSYS-CFX) and the implementation
adopted in this study are considered validated and appropriate
for the investigation of the problem at hand.
4. Results
4.1. Development of the stream-wise velocity prole in the A250 SDiffuser
Fig. 9 shows the variations of the stream-wise velocity prole
along the A250 S-Diffuser at a Re = 80,000. In the AA direction
(Fig. 9a), the velocity prole is symmetrical and gradually attens
downstream, until the centreline velocity becomes smaller than
the velocity in the wall region, with the emergence of an increasingly pronounced overshoot. Along the CC line (Fig. 9b), the symmetrical prole at the inlet gets increasingly distorted towards
the wall at R = 0.5, while the ow near the wall at R = 0.5 is
decelerated. This is an indication of the presence of a secondaryow in the plane of the cross-section as it will be discussed later
in this report.
4.2. Effects of the turbulence intensity at the inlet on the ow and
performance of the A250 S-Diffuser
While rarely investigated in the literature, different operating
conditions could result in signicantly different inlet turbulence

Fig. 9. Development of the stream-wise velocity in the A250 S-Diffuser.

levels, considerably affecting the ow and performance of the diffuser (Sullerey et al., 1983). Three turbulence intensity values are
therefore tested here: Tu = 1%, Tu = 5% and Tu = 10%. The Reynolds
number has been kept constant at 80,000 for all three simulations,
and Figs. 1012 present calculated data at the exit plane of the diffuser (Station 4). The qualitative trends are similar, but increasing
the inlet turbulence intensity (equivalent to enhancing the mixing)
results in a markedly less pronounced overshoot of the velocity
prole in the AA direction (Fig. 10a), and also a less pronounced
distortion of the velocity prole along the CC direction (Fig. 10b).
Fig. 11a shows an initial rapid decrease of the friction coefcients
at the wall at R = 0.5 in the conical section of the A250 S-Diffuser,
followed by a brief increase at the convex bend, and then a gradual
decrease until about X = 0.8 where the boundary layer seems to
detach from the surface though a ow reversal does not actually
occur, corresponding to what is referred to as a boundary layer collision in Talbot and Wong (1982). At the second bend, the friction
coefcient starts increasing again. Near the outer surface of the
rst bend (R = 0.5), the friction coefcients for all three inlet turbulent intensities decrease as the ow passage enlarges, closely
approaching zero, then sharply increase after the concave bend
as a consequence of the ow acceleration in that region of the passage (see also Fig. 9b). This increase is sustained until the convex
bend, at which the friction coefcients drop sharply. The effect of
the inlet turbulence intensity on the quantitative Cf variations is
limited overall, except along the R = 0.5 wall far downstream
where the highest turbulence level correlates with the least friction losses (Fig. 11b). Finally, Fig. 12 shows that the pressure coefcient variations for the three inlet turbulence intensity values
considered are qualitatively similar, indicating a gradual pressure

158

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

0.006
0.4

Tu=1%
Tu=5%
Tu=10%

Tu = 1%
Tu = 5%
Tu = 10%

0.2

Cf

Z*

0.004

-0.2

0.002

-0.4
0

0.5

1.5

U*

0.25

0.5

0.75

0.75

X*

(a) AA line

(a) R* = 0.5
0.4

Tu=1%
Tu=5%
Tu=10%

0.006
Tu = 1%
Tu = 5%
Tu = 10%
0.004

Cf

R*

0.2

-0.2
0.002
-0.4
0

0.5

1.5

U*

(b) CC line
Fig. 10. Effects of the inlet turbulence intensity on the stream-wise velocity prole
at the exit plane of the A250 S-Diffuser.

0.25

0.5

X*

(b) R* = -0.5
Fig. 11. Effects of the inlet turbulence intensity on the calculated friction
coefcients along the wall of the A250 S-Diffuser.

recovery, one of the most important diffuser performance parameters. Quantitatively however, the highest inlet turbulence intensity
returns slightly higher pressure coefcients, while the lowest turbulence intensity value consistently results in lower pressure coefcient values. These results may be explained by the fact that high
turbulence intensity is synonymous to more important mixing, and
are consistent with published experimental observations from the
literature (Sullerey et al., 1983).

4.3. Effects of the Reynolds number on the ow and performance of the


A250 S-Diffuser
Fig. 13 shows the effects of the Reynolds number on the distribution of the stream-wise velocity at the exit plane of the A250 SDiffuser. Four Reynolds number values are considered in this
study: 25,000, 80,000, 116,000 and 209,000 (the latter value being
close to the actual operation condition of the A250 S-Diffuser). The
inlet turbulence intensity is 5% in all cases. At the inlet of the duct,
the proles are identical for all Reynolds numbers, and the difference is also small at the end of the conical section. These proles,
not given here for conciseness, indicate a attening tendency as the
Reynolds number is increased, which is expected. At the exit plane,

the proles along the AA line (Fig. 13a) reveal an increasingly pronounced overshoot in the near wall region, and monotonically
decreasing stream-wise velocities in the centreline when the Reynolds number is increased. Along the CC line (Fig. 13b), the distortion of the velocity prole toward the R = 0.5 wall is also
observed to intensify as the Reynolds number is increased. It is
worth noting that the slope of the velocity prole at about
R = 0.3 suggests that a second overshoot may appear next to the
inner surface of the rst bend should the Reynolds number be further increased. The effect of the Reynolds number on the friction
coefcient variations is illustrated in Fig. 14. The qualitative trends
are again similar for all four Reynolds numbers tested, and the differences in the calculated Cf values are small along the inner surface of the rst bend (slight decrease of the friction coefcients
as the Reynolds number is increased). Along the R = 0.5 wall
however, the friction coefcient variations are extremely sensitive
to the Reynolds number, the simulation at the lowest Reynolds
number returning the highest Cf values over most of the length
of the duct, and most importantly downstream its conical section.
The effect of the Reynolds number on the variations of the pressure
coefcients is shown in Fig. 15, which displays qualitatively
similar results in all cases; however the highest Reynolds number

159

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

0.8
0.4

0.6

0.2

Z*

Cp

25K
0.4

80K

116K
209K

Tu = 1%
-0.2

Tu = 5%
0.2

Tu = 10%
-0.4

0.25

0.5

0.75

0.5

1.5

U*

X*

(a) AA line

(a) R* = 0.5
0.8

0.4
25K
80K
116K

0.2

0.6

Cp

R*

209K
0

0.4
Tu = 1%

-0.2

Tu = 5%
0.2

Tu = 10%
-0.4

0
0

0.25

0.5

0.75

X*

0.5

1.5

U*

(b) CC line

(b) R* = -0.5
Fig. 12. Effects of the inlet turbulence intensity on the calculated pressure
coefcients along the wall of the A250 S-Diffuser.

consistently returned the highest pressure coefcient values on


both walls.
4.4. Sensitivity of the ow and performance parameters to the
geometry of S-shaped diffuser ducts
The seven S-shaped diffuser geometries dened in Figs. 1 and 2,
and an S-shaped duct with a constant cross-sectional diameter and
the same centreline offset as the A250 S-Diffuser are considered in
this paragraph in order to investigate the sensitivity of the hydrodynamic elds and performance indicators to the duct geometry.
Fig. 16 shows the axial variation of the stream-wise velocity distribution in the cross-sectional plane, as well as the variation of the
cross-ow velocity for four of the considered congurations. At
the inlet of the ducts, the stream-wise velocity is symmetrically
distributed, with the highest velocity magnitudes in the centreline
region. The A250 S-Diffuser and the S-Diffuser 1/4 show some
small outward velocity vectors, which is consistent with the fact
that the expansion of their cross-sectional area starts at the inlet
of the ducts. At X = 0.35, both the stream-wise velocity contours
and the cross-ow velocity vectors show a slight movement toward the convex surface of the rst bend. The direction of the secondary-ow vectors however, swiftly reverses direction thereafter,

Fig. 13. Effects of the Reynolds number on the stream-wise velocity prole at the
exit plane of the A250 S-Diffuser.

and the region of the ow where the stream-wise velocity is maximum gradually shifts toward the concave surface under the double effect of centrifugal forces and radial pressure gradients. For the
S-Diffuser 1/4, a separated-ow region is visible at X = 0.81. A separated-ow region is also observed for the S-Diffuser 2/3 at
X = 0.51; it is bigger at X = 0.66, smaller at X = 0.81, and seems
to have disappeared by the exit of the duct. As for the S-Diffuser
3/4, the separated-ow region is rst observed at X = 0.81, and it
is bigger at the exit plane. One single pair of recirculation vortices
is predicted in most cases. Most of these observations are in very
close agreement with those by previous authors on similar geometries (Dean and Hurst, 1959; Patankar et al., 1975; Lee and Yu,
1994). In order to quantify the intensity of the secondary-ow, Table 2 gives the calculated non-uniformity index (NUI) at the exit
plane for the eight geometries considered in this study. The nondiffusing duct (S-Duct) is seen to correspond to the lowest NUI,
agreeing with conclusions in a previous LES study (Abdellatif
et al., 2008). The NUI for the A250 S-Diffuser is relatively high,
but it is the S-Diffuser 2/3 geometry that corresponds to the highest NUI. It is also interesting to note that the normalized maximum
secondary-ow velocity at the exit plane is smaller for all the
geometries in this study than for the RAE2129 in Whitelaw and
Yu (1993). While the most uniform ow possible is desired at
the exit plane of inlet ducts (NUI < 10% qualies as uniform ow)

160

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

0.006

0.8

25K
80K

0.6

116K

0.004

Cp

Cf

209K
25K

0.4

80K
116K

0.002

209K
0.2

0.25

0.5

0.75

X*

0.25

0.5

0.75

X*

(a) R* = 0.5

(a) R* = 0.5

0.006

0.8
25K
80K
116K

0.6

209K

Cp

Cf

0.004

25K

0.4

80K
116K

0.002

209K
0.2

0.25

0.5

0.75

X*

(b) R* = -0.5

0.25

0.5

0.75

X*

(b) R* = -0.5

Fig. 14. Effects of the Reynolds number on the development of the friction
coefcients along the wall of the A250 S-Diffuser.

Fig. 15. Effects of the Reynolds number on the development of the pressure
coefcients along the wall of the A250 S-Diffuser.

(Gopaliya et al., 2011; McDill, 1989; Anderson, 1991), the practical


signicance of the non-uniformity index for a transition duct located between the compressor and the combustion chamber is unknown to us.
Fig. 17 compares the calculated friction coefcients along the
walls of the different variants of the S-shaped diffuser geometries
in this study. Initial friction coefcient curve slopes allow for a
clear differentiation between geometries presenting cross-sectional area expansions starting at the inlet, which show a gradual
decrease of the Cf graphs, and those geometries for which the area
expansion starts farther downstream. The S-shaped duct without
any cross-sectional area expansion is the one that produces the
largest friction losses, which validates the preference for diffusing
ducts in turbomachinery applications. The friction coefcient
graphs also allow for a better appraisal of separated-ow region
features, including nature, location and size. The separated-ow
on the inner surface of the rst bend of the S-Diffuser 1/3
(Fig. 17a) for example qualies as an incipient separation (Lee
and Yu, 1994), that is the friction coefcients decrease down to
zero, without ever changing sign, while the nite negative friction
coefcient values for four of the considered geometries indicate reversed-ow occurrences. The summary in Table 3 shows that the
separated-ow region for the S-Diffuser 2/3 geometry initiates

the earliest, is the biggest, and extends over more than half the
length of the duct, while the ow separates but does not reattach
(within the component) for three other geometrical congurations.
Overall, Fig. 17 and Table 3 suggest that the A250 S-Diffuser is the
most efcient geometry, as it produces the least friction losses, and
it resists ow separation.
The static pressure recovery potentials for each of the geometries considered are shown in Fig. 18. It is seen that the geometries
that present cross-sectional area expansions starting at the inlet of
the duct are the ones that allow for the highest performances.
These geometries also correspond to the shortest lengths of separated-ow region, if any. It is also worth noting that the superior
performance of the A250 S-Diffuser compared to the S-Diffuser
1/4 suggests that the current most common inlet duct design (i.e.
cross-sectional area variation over the entire length of the duct)
is not optimal, at least as far as the static pressure recovery potential is concerned. The S-Diffuser 2/3 and the S-Diffuser 2/4 also
possess decent static pressure recovery potentials, but they are
penalized by the pressure loss between Sections 1 and 2. Limiting
the cross-sectional area expansion to the last section of the duct is
obviously counterproductive. It is also noted that the value of the
static pressure recovery coefcient for the constant cross-sectional
area S-shaped duct is exactly equal to that predicted by the LES

161

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

A250 S-Diffuser

S-Diffuser 1/4

S-Diffuser 2/3

S-Diffuser 3/4

X* = 0

X* = 0.35

X* = 0.51

X* = 0.66

X* = 0.81

X* = 1

Fig. 16. Sensitivity of the stream-wise and cross-ow velocity to the geometry of the duct.

investigation in Abdellatif et al. (2008), which constitutes both an


additional validation and a conrmation that this geometry is not
appropriate for the considered application. Finally, the total pres-

sure loss coefcients are compared in Fig. 19, conrming the better
performance of geometries with cross-sectional area expansions
starting at the inlet. At the other end of the spectrum, the

162

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

Table 2
Sensitivity of the non-uniformity index (NUI) to the geometry of the duct.

0.6

Non-uniformity index

A250 S-Diffuser (S-Diffuser 1/2)


S-Diffuser 1/23/4
S-Diffuser 1/3
S-Diffuser 1/4
S-Diffuser 2/3
S-Diffuser 2/4
S-Diffuser 3/4
S-Duct

0.085
0.065
0.073
0.064
0.112
0.071
0.059
0.027

Pressure Recovery

Duct geometry

0.4

0.2

0.01
A250 S-Diffuser
S-Duct
S-Diffuser 1/2-3/4
S-Diffuser 1/3
S-Diffuser 1/4
S-Diffuser 2/3
S-Diffuser 2/4
S-Diffuser 3/4

0.008

Cf

0.006

-0.2

0.25

0.5

0.75

X*
Fig. 18. Sensitivity of the static pressure recovery to the geometry of the duct (same
legend as in Figs. 17 and 19).

0.004

0.2
A250 S-Diffuser
S-Duct
S-Diffuser 1/2-3/4
S-Diffuser 1/3
S-Diffuser 1/4
S-Diffuser 2/3
S-Diffuser 2/4
S-Diffuser 3/4

0.002
0.15

0.25

0.5

0.75

CL

0.1

X*

(a) R* = 0.5
0.05
A250 S-Diffuser
S-Duct
S-Diffuser 1/2-3/4
S-Diffuser 1/3
S-Diffuser 1/4
S-Diffuser 2/3
S-Diffuser 2/4
S-Diffuser 3/4

0.008

0.25

0.5

0.75

X*
Fig. 19. Sensitivity of the total pressure loss coefcient to the geometry of the duct.

Cf

0.006

0.004

S-Diffuser 2/3 and the S-Diffuser 3/4 produce total pressure losses
equivalent to those of the constant cross-sectional area, S-shaped
duct.

0.002

5. Conclusions
0

0.25

0.5

0.75

X*

(b) R* = -0.5
Fig. 17. Sensitivity of the friction coefcients to the geometry of the duct.
Table 3
Separated ow regions for the S-shaped diffuser geometries investigated.
Separation Reattachment Separated
point
point
ow
length
A250 S-Diffuser (S-Diffuser 1/2)
S-Diffuser 1/23/4
S-Diffuser 1/3
S-Diffuser 1/4
S-Diffuser 2/3
S-Diffuser 2/4
S-Diffuser 3/4
S-Duct

n/a
0.765
0.569
0.747
0.443
0.618
0.716
n/a

n/a
n/a
0.744
0.989
0.969
n/a
n/a
n/a

n/a
n/a
0.175
0.242
0.526
n/a
n/a
n/a

An investigation has been conducted on the Allison 250 gas turbine S-shaped diffuser which connects the exit of the compressor
to the inlet of the combustion chamber using the standard je
and jx models, as well as the Shear-Stress Transport (SST) model
implemented in the commercial code ANSYS-CFX. Data measured
using an in-house purpose-designed test rig and other experimental data taken from the literature were used for validation. The
standard jx model was shown to qualitatively reproduce the
general trends, but the quantitative agreement with the experimental data was poor. The standard je model performed even
worse. Conversely, the SST model successfully passed the validation tests and was therefore adopted for the remainder of the analysis. Among other calculated results, the stream-wise velocity
prole was observed to atten in the initial divergent section,
experience a slight shift toward the inner surface at the rst bend,
before a gradually intensifying secondary-ow is initiated, pushing
the region of the ow with the highest velocity toward the concave
surface of the rst bend. This was accompanied by an increase of

G.G. Lee et al. / International Journal of Heat and Fluid Flow 42 (2013) 151163

the friction coefcient in that region. In the direction normal to the


offset, the velocity distribution remained symmetrical, with two
overshoot regions close to the wall. The distortion of the streamwise velocity intensied when the inlet turbulence intensity was
decreased, with a small effect on the friction and pressure coefcients. On the other hand, the severity of the distortion of the
stream-wise velocity prole was shown to increase with the Reynolds number. The static pressure recovery potential also increased
with the Reynolds number, while higher Reynolds numbers consistently translated into lower wall friction coefcients. Six variations
of the diffuser geometry have been considered in this study, all
having the same total cross-sectional area ratio and centreline offset, but differing on the location and length of the area expansion
section. The general variation trends were the same as those of the
Allison 250 diffuser, but the decrease of the friction coefcients
near the inner surface of the rst bend continued beyond the zero
line, and separated-ow regions (and reversed-ow regions in
some cases) of different sizes and at different locations were observed for all of these geometries. The friction coefcients were
the lowest overall, the static pressure recovery was the highest,
and the total pressure loss coefcient was among the lowest for
the base geometry. Particularly worth noting is that the Allison
250 S-shaped diffuser duct has overall better performance characteristics than its variant presenting a cross-sectional area expansion extending over its entire length. This is interesting because
most common inlet ducts actually have cross-sectional area expansion over their entire length.
Acknowledgements
The nancial support of Standard Aero Ltd. (SAL), the National
Sciences and Engineering Research Council of Canada (NSERC)
and the Royal Military College Academic Research Program (ARP)
is acknowledged.
References
Abdellatif, O.E., Abd Rabbo, M., Abd Elganny, M., Shahin, I., 2008. Area ratio effect on
the turbulent ow through a diffusing S-duct using large-eddy simulation. In:
AIAA 6th Int. Energy Conversion Engineering Conf., Cleveland, OH, Paper AIAA2008-5726.
Abe, K., Ohtsuka, T., 2010. An investigation of LES and hybrid LES/RANS models for
predicting 3-D diffuser ow. Int. J. Heat Fluid Flow 31, 833844.
Anand, R.B., Rai, L., Singh, S.N., 2003. Effect of the turning angle on the ow and
performance characteristics of long S-shaped circular diffusers. Proc. Inst. Mech.
Eng. Part G: J. Aerospace Eng. 217, 2941.
Anderson, B.H, 1991. The aerodynamic characteristics of vortex ingestion for the F/
A-18 inlet duct. In: AIAA 29th Aerospace Sciences Meeting, Reno, NV, Paper
AIAA-91-0130.
Azad, R.S., 1996. Turbulent ow in a conical diffuser: a review. Exp. Therm. Fluid Sci.
13, 318337.
Barth, T.J., Jespersen, D.C., 1989. The design and application of upwind schemes on
unstructured meshes. AIAA J., 198366.

163

Cherry, E.M., Elkins, C.J., Eaton, J.K., 2008. Geometric sensitivity of threedimensional separated ows. Int. J. Heat Fluid Flow 29, 803811.
Dean, W.R., Hurst, J.M., 1959. Note on the motion of uid in a curved pipe.
Mathematika 6, 7785.
El-Behery, S.M., Hamed, M.H., 2011. A comparative study of turbulence models
performance for separating ow in a planar asymmetric diffuser. Comput.
Fluids 44, 248257.
Gopaliya, M.K., Kumar, M., Kumar, S., Gopaliya, S.M., 2007. Analysis of performance
characteristics of S-shaped diffuser with offset. Aerospace Sci. Technol. 11, 130
135.
Gopaliya, M.K., Goel, P., Prashar, S., Dutt, A., 2011. CFD analysis of performance
characteristics of S-shaped diffusers with combined horizontal and vertical
offsets. Comput. Fluids 40, 280290.
Gullman-Strand, J., Tornblom, O., Lindgren, B., Amberg, G., Johansson, A.V., 2004.
Numerical and experimental study of separated ow in a plane asymmetric
diffuser. Int. J. Heat Fluid Flow 25, 451460.
Herbst, A.H., Schlatter, P., Henningson, D.S., 2007. Simulations of turbulent ow in a
plane asymmetric diffuser. Flow Turbul. Combust. 79, 275306.
Jakirlic, S., Kadavelil, G., Kornhaas, M., Shafer, M., Sternel, D.C., 2010. Tropea:
numerical and physical aspects in LES and hybrid LES/RANS of turbulent ow
separation in a 3-D diffuser. Int. J. Heat Fluid Flow 31, 820832.
Lee, K.M., Yu, S.C.M., 1994. Computational studies of ows in the RAE 2129 S-shaped
diffusing duct. In: 32nd AIAA Aerospace Sciences Conf., Reno, NV, Paper AIAA940658.
Lee, G.G.W., Allan, W.D.E., Goni Boulama, K., 2012. Numerical and experimental
analysis of the airow inside an A250 diffuser tube. In: Int. Gas Turbine Institute
Conf., Copenhagen, Denmark, Paper GT201269708.
McDill, P.L., 1989. An Experimental Evaluation of S-Duct Inlet-Diffuser
Congurations for Turboprop Offset Gearbox Applications. NASA, Report,
NASA-CR-179454.
Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for engineering
applications. AIAA J. 32, 15981605.
Ohlsson, J., Schlatter, P., Fischer, P.F., Henningson, D.S., 2010. Direct numerical
simulation of separated ow in a three dimensional diffuser. J. Fluid Mech. 650,
307318.
Patankar, S.V., Pratap, V.S., Spalding, D.B., 1975. Prediction of turbulent ow in
curved pipes. J. Fluid Mech. 67, 583595.
Rhie, C.M., Chow, W.L., 1983. Numerical study of the turbulent ow past an airfoil
with trailing edge separation. AIAA J. 21, 15251532.
Rudolf, P., Desova, M., 2007. Flow characteristics of curved ducts. Appl. Comput.
Mech. 1, 255264.
Saha, K., Singh, S.N., Seshadri, V., 2007. Computational analysis on ow through
transition S-diffusers: effect of inlet shape. J. Aircraft 44, 187193.
Schneider, H., von Terzi, D., Bauer, H.J., Rodi, W., 2010. Reliable and accurate
prediction of three-dimensional separation in asymmetric diffusers using
Large-Eddy Simulation. J. Fluids Eng. 132, 031101-1031101-8.
Sparrow, E.M., Abraham, J.P., Minkowycz, W.J., 2009. Flow separation in a diverging
conical duct: effect of Reynolds number and divergence angle. Int. J. Heat Mass
Transfer 52, 30793083.
Sullerey, R.K., Chandra, B., Muralidhar, V., 1983. Performance comparison of straight
and curved diffusers. Def. Sci. J. 33, 195203.
Talbot, L., Wong, S.J., 1982. A note on boundary-layer collision in a curved pipe. J.
Fluid Mech. 122, 505510.
Towne, C.E., 1984. Computation of viscous ow in curved ducts and comparison
with experimental data. In: AIAA Aerospace Sc. Meeting, Reno, NV, Paper AIAA84-0531.
Whitelaw, J.H., Yu, S.C.M., 1993. Turbulent ow characteristics in an S-shaped
diffusing duct. Flow Meas. Instrum. 4, 171179.
Wu, X., Schluter, J., Moin, P., Pitsch, H., Iaccarino, G., Ham, F., 2006. Computational
study on the internal layer in a diffuser. J. Fluid Mech. 550, 391412.
Yang, Y.T., Hou, C.F., 1999. Numerical calculation of turbulent ow in symmetric
two-dimensional diffusers. Acta Mech. 137, 4354.

You might also like