You are on page 1of 10

Journal of Thermal Science Vol.19, No.

4 (2010) 300309

DOI: 10.1007/s11630-010-0387-8

Article ID: 1003-2169(2010)04-0300-10

Application and Comparison of SST Model in Numerical Simulation of the Axial


Compressors
YIN Song1, JIN Donghai2, GUI Xingmin3 and ZHU Fang
Aero-Engines Simulation Research Center, School of Jet Propulsion, Beijing University of Aeronautics and Astronautics, Beijing, 100191, China
Science Press and Institute of Engineering Thermophysics, CAS and Springer-Verlag Berlin Heidelberg 2010

The shear-stress transport (SST) turbulence model is incorporated into Navier-Stokes equations to simulate a turbomachinery flowfield. A staggered finite volume method is used to make the mean flow equations and turbulence model equations strongly coupled and enhance the stability of the numerical computation. The implicit
treatment of the source terms is applied to the SST model. A steady state solution is obtained using five-stage
Runge-Kutta time-stepping scheme with local time stepping and residual smoothing to accelerate convergence.
The wall distance d, a key parameter in the SST model, is solved by a partial differential equation. The validations
of the code are conducted on rotor 37, wp11 at design and off-design conditions by comparison with measurements and the Spalart-Allmaras (SA) turbulence model. The flow within the tip is calculated with a multi-block
grid.

Keywords: SST model, staggered finite volume method, turbomachinery, simulation, multi-block grid

Introduction
Of all the flowfields in the current engineering configurations, the flow within axial-flow turbomachines is
probably one of the most complicated. The prediction of
complex turbulence flow is a nontrivial problem. Since
the early 1980s when fully three-dimensional methods
first became available, CFD has been a useful and important tool to investigate the complex flow phenomena
and been applied in the design system. Especially with
the fast development of powerful computer and efficient
numerical methods, CFD becomes more and more practical, desirable and absolutely necessary for aerodynamic
problems. As widely recognized in the turbomachinery
aerodynamics CFD community, the current generation of
CFD codes provides very useful numerical solutions for
design purpose. Thus, the fidelity of CFD simulations is

crucial and one of the important issues is the turbulence


model. The current state of the art in turbulence models
for turbomachinery simulation is mainly the systematic
use of two-equation models, which offer a good balance
between accuracy, robustness and cost effectiveness for
engineering problems.
The notable shear stress transport (SST) turbulence
model was proposed by Menter[1]which is a zonal twoequation turbulence model that is the k- model near the
wall and gradually transited to the k- model away from
the wall. It has the advantage that it combines the superior performance of the k- model in the near wall regions and the freestream independence of the k- model
in the farfield. In recent years, it has widely and successfully made its way into most industrial aerodynamic predictions. It has better performance over other two-equation models especially for adverse pressure-gradient and

Received: April 2010


GUI Xingmin: Professor
www.springerlink.com

YIN Song et al. Application and Comparison of SST Model in Numerical Simulation of the Axial Compressors

separated flows. From this point, this model appears very


attractive and favorable in compressor industry. In this
paper, the SST turbulence model is employed to calculate
the eddy viscosity.
The finite-volume method with a multi-stage timestepping scheme proposed by Jameson [2] has the advantage of separated spatial and time discretizations and is
easy to implement. Much attention should be paid to the
solution of the turbulence model equations, particularly
their coupling with the Navier-Stokes equations [3, 4, 5].
For the Navier-Stokes equations strongly coupled with
turbulence model in the numerical scheme, the staggered
control volume method [6] in space is applied in this paper.
With implicit residual smoothing, locally varying time
steps and suitable implicit treatment of the source terms
of the SST model, excellent convergence properties are
obtained.
In order to investigate numerical characteristics of the
three-dimensional compressible Navier-Stokes solver, the
calculated flowfields of two types of transonic compressor rotors with SST model are compared with experimental data and the Spalart-Allmaras (SA) turbulence
model. One is NASA Rotor 37, the other is WP11. The
flow within the tip is calculated with a multi-block grid.
The main purpose of this paper is to investigate the performance of SST turbulence model for turbomachinery
flows especially with tip leakage under different operation conditions.

Calculation Method
Governing Equations
The 3-D compressible, Reynolds-averaged Navier-Stokes equations in an integral conservation form in
Cartesian coordinates are expressed as:

UdV + v ( Fi + Gj + Hk ) dS =
t

v ( Fv i + G v j + H v k ) dS + TdV

(1)

here, U= [ , u, v, w, E ] are the five conservation


1

variables; F, G and H are the convective terms; FV, GV


and HV represent the viscous shear stress and heat flux
terms; T are the additional terms from centripetal force
and Coriolis force by coordinate transformation; t is the
time step; V is the control volume; S is the surface area
vector of the cell; u, v and w are the components of the
relative velocity in the x, y, and z directions respectively;
is the density; E is the total energy.
Turbulence Model
The SST model was originally proposed by Menter[1]
to combine the best elements of the k-, the k- and the

301

JK models. Subsequently, several modifications are carried out to gain the better robustness [7, 8].
The modified SST turbulence model is given as:

k
t

+ ui

k
xi

= ij
+

+ ui

xi

ui
x j

k
( + k t )

x j
x j

ij i 2
x j
t

(2)


( + t )

x j
x j

+2 (1 F1 ) 2

1 k

x j x j

The turbulent eddy viscousity is defined as follows:


t =

a1k

(3)

max ( a1 ; SF2 )

where S is the invariant measure of the strain rate instead


of vorticity in original formulation; F1 and F2 are blending functions that ensure the proper selection of k- and
k- zones without users interference. The SST limitation
on t based on Bradshaws assumption that the shear
stress in a boundary layer is proportional to turbulent
kinetic energy significantly improves the model performance for adverse pressure gradient boundary layers.
The production term Pk is limited in the SST model to
improve the numerical robustness of the algorithm by
eliminating the unrealistic buildup of eddy viscosity in
the stagnation regions and removing the occurrence of
spikes in the eddy viscosity due to numerical wiggles.
The modification is the use of the factor 10 not 20 in
original version:
Pk = min( Pk ,10 Dk )
(4)
During the process of the iteration, the most significant numerical problem is the appearance of nonphysical
negative values of turbulence variables. Various limiters
have been devised to deal with this problem. To ensure
the positive of turbulence variables, limiter on k proposed
by Ilinca [9] is adopted. If k is too small, it is replaced by
kmax/dk.Here, kmax is the maximum of k within the calculation domain and dk is the constant determined by users.
Meanwhile, the lower bounds for the specific dissipation
are limited according to Zheng and Liu [10]:

3
2

1 uk

3 xk

2 Sij

ui

ij

xi

(5)

Flow Solver
A central-difference with additional second and fourth
artificial dissipation scheme is used for the convective

302

J. Therm. Sci., Vol.19, No.4, 2010

terms in the framework of finite volume method. As for


the viscous terms, a conventional central-difference
scheme is evaluated. Much attention should be paid to
the implementation of turbulence model equations into
Navier-Stokes equations. Liu[6] developed a staggered
finite volume scheme for discretizing the Navier-Stokes
and k- turbulence model equations to achieve strong
coupling of the two sets of equations and avoid the possibility of an odd-and-even decoupling, by solving the
Navier-Stokes equations with a cell-centered scheme,
whereas the turbulence model equations with a cell-vertex scheme. This paper extends this method to the SST
model.
Figure 1 illustrates the staggered control volume
scheme, in which Wi is the Favre-averaged velocity vector, xj is the position vector. Both the flow variables ,
Wi, E and the derivatives of the turbulence variables
k/xj, /xj are stored at the cell centers marked by the
circles. The turbulence variables k, and the derivatives
of the velocity vector Wi and temperature T, Wi/xj,
T/xj are stored at the cell vertexes. The fluxes in the
Navier-Stokes equations are estimated over the four cell
faces of the control cell while the auxiliary control
volume is used to integrate the SST model.

employed, with evaluations of the dissipation and the


viscous terms only at the first, third, and fifth stage. Implicit residual smoothing and locally varying time steps
are used to accelerate convergence with application to
both the Navier-Stokes and the turbulence model equations.
The Treatment of the Source Terms
One of the important aspects in the discretization of
the two-equation turbulence model in a separate subroutine decoupled from the mean flow equations is the implicit treatment of the source terms. Proper evaluation of
the source terms is very important to the stability of the
computation. According to equation (2), the semi-discrete
SST equations can be written as:
( k )
t
( )

+ Rk = 0

(7)
+ R = 0

here, Rk and R are the residual for the k and , respectively.


Thus, the explicit time-marching solution for the SST
model equations can be written as:
( k

n +1

k )

+ Rk = 0

t
(

n +1

(8)

+ R = 0

Linearizing the production, the destruction and the additional cross diffusion terms, we have
( k

n +1

k )

t
(

n +1

Figure 1

Schematic of staggered control volume scheme

A system of ordinary differential equations in time can


be obtained after discetizing equation (1) and (2) in space:
Thus, we obtained
dU N S 1
d t + (Qc Qv D ) T = 0
(6)

dU SST + 1 (Q Q D) T = 0
c
v

d t
where Qc, Qv are the net convective and diffusive fluxes
out of the control cell and D is the additional artificial
dissipative term, respectively.
A five-stage Runge-Kutta time stepping scheme is

+ Rk = ( k
n

n +1

+ R = (

k )

n +1

( Pk Dk )

( P D + CDk )

(9)
where P, D, CD are the production, the destruction and
the additional cross diffusion terms respectively.
The right-hand side of equation (9) can be substituted
by the approximate expressions [1]:

( Pk Dk )

2 Dk

min(0, Pk )

( P D + CDk )

2 D + CDk (1 F1 )

(10)

Then, we obtained
k n +1 k n =

Rk t

2 Dk min(0, Pk )

k
k

1 + t

n +1

R t

2 D + CDk (1 F1 )

1 + t

(11)

YIN Song et al. Application and Comparison of SST Model in Numerical Simulation of the Axial Compressors

Grid System
The grid used in the solver is a simple H mesh via
stacking the 2-D streamface grid in the spanwise direction. The streamface grid is obtained by solving the elliptical equations. For good orthogonality and displacements of grids near walls, the source terms of the elliptical equations are adjusted according to the method of
Hilgenstock[11]. To satisfy the requirement of the distance
of the first point from the wall for the turbulence model,
the contour line of the wall distance d required by turbulence model around the wall boundary is first created,
just like the highlight line in Figure 2.
The tip clearance flow is described with the patched
grid technique, which connects the point on the pressure
side and the corresponding one on the suction side and
offers high grid resolution. Meanwhile, the distribution
along the pitchwise direction is consistent with the main
block. It should be noted that the grids at the leading and
trailing edge in the tip clearance block are triangularprisms not hexahedrons as shown in Figure 3. Therefore,

303

additional treatment is adopted when discretizing the


equations and the axial artificial dissipation is not included.
Computations of Wall Distances
The normal wall distance to the nearest wall d is a key
parameter in SST model, included in the blending functions. It is gotten by the solution of a partial differential
equation[12,13, 14], which can be combined with turbulence
models and can get more exact normal distance than calculating the minimum distance between discrete points
on the wall and the points in the domain if the grid lines
are not orthogonal to the surface. The partial differential
equation used in this paper is based on the equation derived by Fares [14].
The partial differential equation explicitly advanced in
an artificial time with cr being a relaxation factor
reads:
G t = cr [(G G x ) x + (G G y ) y + (G G z ) z
+( 1) G (G xx + G yy + G zz ) - G 4 ]

wit

0.2
= (1 + 2 )

(12)

1
1

G G0
where G0 is the value of the inverse of wall distance d on
the wall.
The initial and boundary conditions, as well as the detailed derivation of the equation, are described in Refs.14.
The proposed formulation has computational advantages
and can be favorably incorporated into the governing
equations. For conveniently utilizing the above numerical
methods of the Navier-Stokes and the turbulence model
equations and conjunction with them, the equation (12)
above is written in integral form as:
d=

Leading edge

Figure 2

Trailing edge

mesh at the leading edge and trailing edge

G(

n+1)

n
- G( )

dV = v (GG i ) dS - G 4 dV

(13)

According to the form of equation (1), the same form


is given by:
GG , G = GG , H = GG , T = G 4
U = [ G ] , Fv =

x v y v z

Results and Discussion

Figure 3 mesh at tip clearance

NASA Rotor 37
The NASA Rotor 37 is a well-known turbomahchinery
test case. This rotor has 36 blades, nominal speed
17188.7 r/min, and a design pressure ratio of 2.106 at a
mass flow of 20.19 kgs1.This test case has been computed by numerous researchers. Various theoretical pre-

304

J. Therm. Sci., Vol.19, No.4, 2010

dictions from different codes [15-19] and detailed experimental data [20, 21] will be very useful in calibrating numerical tools. Therefore, Rotor 37 is preferred as a test
case to assess the predictive capabilities of turbomachinery CFD tools.
Computational results obtained with the SST model
are compared with the measurements for various operating points, using multi-block grids with 49 nodes in the
blade-to-blade direction, 49 nodes in the spanwise direction, 129 nodes in the streamwise direction within the
domain, and 65 axial stations,10 pitchwise stations, 10
radial stations within the tip-clearance gap.
The overall rotor performance maps at the 60, 80, and
100 percent design speed are shown in Figure 4. The
computed choke mass flow was 21.002 kgs1, while the
measured was 20.930.14kgs1 . At the same speed, as
the flow reduces, the efficiency is gradually close to the
observed in the experiment. This tendency was explained
by Arima[16] that may result from the inlet boundary. The
deviation of the calculated value from the experimental
data shown in Figure 4 in terms of the total pressure ratio
at the off-design speeds is a little large. Some possible
causes of the deviation will be discussed later.
The calculated flow fields at the 30,50,70,90 percent
span at 98% choke mass flow at the design speed are
analyzed and compared with the experimental data, as
shown in Figure 5. The results show that the calculated
flow fields are in good agreement with the measured in
terms of the location and strength of the passage shock
structure. But the blade wake is a little thinner than the
experiment.
Figure 6 and Figure 7 compare the relative Mach
number distributions in tip clearance flow between numerical results and test data at design and off-design
speed. As shown in Figure 6, for 98 percent mass flow
rate, the passage shock in the tip is distorted due to the
tip clearance vortex and low Mach number region is

formed after the passage shock. The location and area of


predicted low Mach number region is very close to the
experiment. Meanwhile, the front passage shock moves
forward at the interaction location and bow shock is attached to the leading edge. The passages shock disappears when the blade loading reduces at 60 percent rotor
speed as shown in Figure 7.
In comparison with the measurements, prediction of
spanwise distributions of total pressure and total temperature is presented in Figure 8 for conditions of 98 and
92.5 percent choke mass flow rate. We used station 4 as
the previous researchers did. At the 92.5 percent choke
flow rate, the calculated property distribution agrees better with the measurements of 98 percent choke mass rate.
This observation suggests that the calculated loss is
greater than the reality and this will result in the lower
efficiency as seen in Figure 4. At off-design speed, the
calculated data is smaller than the measurements. It indicates that the results of the lower calculated flow may
agree better with the measurements. The computation
underestimates total pressure over the entire span at a
given mass flow in contrast to overestimating the value
by other researchers, especially in the vicinity of tip
clearance flow. This may explain that the calculated efficiency agrees better with the measurements than other
researcher, presumably associated with the SST model
capability of predicting the separated and shear flow better than other turbulence models. It is noteworthy that a
dip in the spanwise distribution of total pressure near the
hub at the design speed appears which is explained by
Hah [19] that is due to the vortex formed at the corner of
the hub and the blade suction surface toward the rear of
the blade passage. Later on, Hah testified this point
through the LES [19].
The results from comparison between calculations and
measurements agree well with other researchers, which is
that the absolute values of properties are not predicted

2.2

0.98
0.96

Design Speed

1.8
1.6
80% Speed
1.4
1.2
60% Speed
1.0
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
Mass Flow Rate/Mass FLow Rate at Choke for Design Speed

Adiabatic Efficiency

Total Pressure Ratio

2.0

experiment
SST
SA

0.94

experiment 60% Speed


SST
SA

80% Speed

0.92
0.90

Design Speed

0.88
0.86
0.84
0.82
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
Mass Flow Rate/Mass FLow Rate at Choke for Design Speed

Figure 4 Rotor Performance for NASA Rotor 37 at 60, 80 and 100 percent of the design speed

YIN Song et al. Application and Comparison of SST Model in Numerical Simulation of the Axial Compressors

Figure 5

305

Calculated and experimental relative Mach number contours at 98 percent choke mass rate for Rotor 37

Figure 6 Calculated and experimental relative Mach number contours at 95% span and design speed

well, but the shape of properties distribution is similar to


the experimental data. However, Hah pointed out that
accurate prediction of the shapes of property distributions could be more useful than absolute values of integrated properties.
WP11
To validate the code further, another axial transonic

compressor rotor WP11 was studied. The Rotor was experimentally investigated by Gui[22,23]. The basic specifications of the rotor are indicated in Table 1. The calculation domain consists of the region between station 1 and
station 2 as shown in Figure 10, corresponding to the
measuring position in the experiment. Figure 11 shows a
three-dimensional view of the grids on the blade surface
and hub endwall. The computational grids consist of 49

306

J. Therm. Sci., Vol.19, No.4, 2010

100

100

80

80

60

60

% Span

% span

Figure 7 Calculated and experimental relative Mach number contours at 95 percent span and 60 percent rotor speed

40

40
20

20
0
1.8

1.9

2.1

2.2

0
1.24

2.3

Total Pressure Ratio

1.3

experiment 98% flow

SST 98% flow

SA 98% flow

experiment 92.5% flow

SST92.5%flow

SA92.5%flow

100

80

80

60

60

% Span

100

40
experiment
SST
SA

20

1.32

1.34

1.0

1.1

1.2

40
20

0
1.3

1.4

Total Presuure Ratio

Figure 9

1.28

Total Temperature Ratio

Comparison of total pressure and total temperature distribution at the design speed

% Span

Figure 8

1.26

0
1.00

experiment
SST
SA
1.03

1.06

1.09

1.12

Total Temperature Ratio

Comparison of total pressure and total temperature distribution at the 60 percent speed

nodes in the blade-to-blade direction, 49 nodes in the


spanwise direction and 113 nodes in the streamwise direction. To describe the clearance flow, 11 nodes are distributed tangentially across the blade tip, 9 nodes in radial direction between the blade tip and shroud, and 65

nodes along the chord.


The calculations were carried out for three operating
speed conditions (100, 90, and 80 percent of the design
speed). Figure 12 shows the overall rotor performance
maps at different speeds. A few discrepancies remain in

YIN Song et al. Application and Comparison of SST Model in Numerical Simulation of the Axial Compressors
Table 1 Basic specifications of WP11
Number of Rotor Blades

17

Inlet Shroud Diameter

[mm]

355.8

Midspan Chord

[mm]

80.5

Aspect Ratio

0.956

Hub/Tip Radius Ratio

0.565

Blade Solidity at Tip

1.24

Mass Flow

[kg/s]

Pressure Ratio

1.6

Inlet Tip Relative Mach Number

1.38

Inlet Hub Relative Mach Number

0.963
[r/min]

Rotor Tip Speed

[m/s]

Tip Clearance at Design Speed

[mm]

the calculated loss in this region is lower than the experiment. Meanwhile, the computational passage shock is
weaker than that captured by the experiment. Thus, the
loss from the passage shock may be underestimated.
Based on the comprehensive consideration of the impact
of the underestimated loss of the two respects, the calculated efficiency is much higher than the experiment.

Conclusions

13.5

Rotational Speed

In the present work, the shear-stress transport model


was extended to turbomachinery applications. Threedimensional flow fields for two types of transonic com-

22000
409.85
0.853

pressure ratio characteristics, but considerable discrepancies exist in the efficiency characteristics, especially under the off-design conditions.
Figure 13 shows the static pressure contours under the
conditions of the symbol A, B and C in Figure 12. The
calculated results are presented on the left while the observation is presented on the right. An evident detached
bow shock exists in different operating conditions and
interacts with tip clearance flow. As the mass flow decreases, the position of the detached bow shock is apparently pushed forward and the size of the expansion region is reduced. The reacceleration region is reduced and
almost disappears near stall. The leakage flow at the
leading edge grows much stronger with the increase of
the incidence angle, which results in an evident low
pressure region and interacts with the passage shock.
Influenced by the leakage flow at the leading edge, the
passage shock is nearly normal to the suction surface.
The interaction between the leakage flow and the passage
shock by calculation is more evident than the measurements. But the area of low pressure formed by the leakage flow is smaller than the experiment, which is to say,

Figure 10 Meridional view of WP11

Figure 11

1.8

1.6
Design Speed
1.5
C
B

1.4

1.3

90%Speed
10

11
12
Mass Flow

13

0.92
0.88
0.84
90% Speed

0.80
0.76

80% Speed
9

experiment
SST
SA

0.96

Adiabatic Efficiency

Total Pressure Ratio

Computational grid for WP11

1.00

experiment
SST
SA

1.7

1.2

307

14

0.72

Design Speed

80% Speed
9

10

Figure 12 Rotor Performance for WP11 at 80, 90 and 100 percent of the design speed

11
12
Mass Flow

13

14

308

Figure 13

J. Therm. Sci., Vol.19, No.4, 2010

Pressure contours at the tip for 90 percent of the design speed

pressor rotors (NASA rotor 37, WP11) were used for


validation of the CFD code. The details of tip-clearance
flow were described with a multi-block grid. The computation results accord with the relevant experimental
data qualitatively. The total pressure deficit near the hub
at the rotor exit is calculated correctly. However, quantitative discrepancies remain in predicting the overall performance and the spanwise distributions of the aerodynamic parameters as other researchers did. The calculation successfully describes the tip clearance flow and its
interaction with the main flow. Especially the location
and area of the low Mach number region after the passage shock is very close to the experiment, but the loss of
interactions between the clearance flow and the shock is
underestimated, which results in the efficiency overestimation as compared with the experiment. Therefore, numerical procedures need further refinements to enhance
the applicability of the code to various flowfields.
In comparison with SA computations, SST model
shows improvement to some extent in terms of the overall performance and the distributions of the aerodynamic
parameters. For combining the best elements of the k-,
the k- and the JK models, SST may enhance the quantitative prediction of the complex turbulent flows with
strong adverse pressure-gradient and separated flows
encountered in axial compressors. Efforts will be made to
conduct more detailed investigations of the SST model to
develop its capability of simulating turbomachinery.

Acknowledgements
We would like to acknowledge the fact that this work
was performed under the Aero-Engines Simulation Research Center in Beijing University of Aeronautics and
Astronautics. In particular, it is also worth underlining
that the project was supported by the National Natural
Science Foundation of China under Contract 50676004
and 50736007.

References
[1] Menter, F. R., Zonal two equation k- turbulence models
for aerodynamic flows, the 24th Fluid Dynamics Conference, July 6-9, 1993,Orlando, Florida, AIAA-93-2906.
[2] Jameson, A., Schmidt W., Turkel E., Numerical solutions
of the Euler equations by finite volume methods with
Runge-Kutta time stepping schemes, the 14th Fluid and
Plasma Dynamics Conference, June 2325, 1981,Palo
Alto, California, AIAA-81-1259.
[3] Turner, M. G., and Jennions, I. K., An investigation of
turbulence modelling in transonic fans including a novel
implementation of an implicit k-epsilon turbulence model,
International Gas Turbine and Aeroengine Congress and
Exposition, 37th, Cologne, Germany; UNITED STATES;
1-4 June 1992. ASME 92-GT-308
[4] Kunz, R. F., and Lakshminarayana, B., Explicit Navier-Stokes computation of cascade flows using the
k-epsilon turbulence model, AIAA Journal, Vol. 30, No.

YIN Song et al. Application and Comparison of SST Model in Numerical Simulation of the Axial Compressors
1, 1992, pp. 1322.
[5] Minjiang Li. Numerical investigation and design of transonic compressors with tip clearance flow using multiblock mesh, A Dissertation for Ph.D. Degree, Dept of Jet
Propulsion, Beijing University of Aeronautics and Astronautics, Beijing, 2004
[6] Liu, F., and Zheng, X., Staggered finite volume scheme
for solving cascade flow with a k- turbulence model,
AIAA Journal, Vol. 32, No. 8, 1994, pp. 15891579.
[7] Menter, F. R., Kuntz M., Langtry R., Ten years of industial experience with the SST turbulence model, Turbulence, Heat and Mass Transfer 4, ed: K. K. Hanjalic, Y.
Nagano, and M. Tummers, Begell House, Inc., 2003, pp.
625632.
[8] Menter, F. R., Review of the Shear-Stress Transport turbulence experience from an industrial perspective, International Journal of Computational Fluid Dynamics, Vol.
23, No. 4, 2009, pp. 305316.
[9] Ilinca, F., and Pelletier, D., Positivity preservation and
adaptive solution for the k- model of turbulence, AIAA
Journal, Vol. 36, No. 1, 1998, pp. 4450.
[10] Zheng, X., and Liu, F., Staggered upwind method for
solving Navier-Stokes equations with k- turbulence
model, AIAA Journal, Vol. 33, No. 6, 1995, pp. 991998.
[11] Hilgenstock A., A fast method for the elliptic generation
of three dimensional grids with full boundary control,
Num. Grid Generation in CFM88, Pineridge Press LD.,
1988
[12] Tucker, P. G., Assessment of geometric multilevel convergence robustness and a wall distance method for flows
with multiple internal boundaries, Applied Mathematical
Modeling, Vol. 22, 1998, pp. 293311
[13] Tucker, P. G., Rumsey, C. L., et al., Computations of
wall distances based on differential equations, AIAA
Journal, Vol. 43, No. 3, 2005, pp. 539549.
[14] Fares E., Schrder W., A differential equation for ap-

[15]

[16]

[17]

[18]

[19]
[20]

[21]

[22]

[23]

309

proximate wall distance, International Journal for Numerical Methods in Fluids, Vol. 39, 2002, pp. 743762
Hah, C., Loellbach, J., Development of hub corner stall
and its influence on the performance of axial compressor
blade rows, ASME Journal of Turbomachinery, Vol. 121,
1999, pp. 6777.
Arima, T., Sonoda, T., Shirotori, M., Tamura, A., and Kikuchi, K., A numerical investigation of transonic axial
compressor rotor flow using a low-Reynolds-Number k-
turbulence model, ASME Journal of Turbomachinery, Vol.
121, 1999, pp. 4458.
Gerolymos, G.A., and Vallet, I., Wall-normal-free Reynolds-Stress model for rotating flows applied to turbomachinery, AIAA Journal, Vol. 40, No. 2, February 2002,
pp. 15891579.
Chima, R. V., Calculation of tip clearance effects in a
transonic compressor rotor, ASME Journal of Turbomachinery, Vol. 120, 1998, pp. 218229.
Hah, C., Large eddy simulation of transonic flow field in
NASA Rotor 37, NASA-TM-2009-21562
Suder, K.L., Experimental investigation of the flow field
in a transonic, axial flow compressor with respect to the
development of blockage and loss, NASA TM 107310,
1996.
Suder, K.L., and Celestina, M. L., Experimental and
computational investigation of the tip clearance flow in a
transonic, axial flow compressor rotor, ASME Journal
of Turbomachinery, Vol. 118, No. 2, 1996, pp. 218229.
Gui, X.M. Wang, T. Q., Yu, Q., Li, Y.M., Investigation on
the tip clearance flow field in a transonic compressor rotor, Journal of Engineering Thermophysics, Vol. 19, No.
5, 1998, pp. 553558.
Gui, X.M. Nie, Ch.Q., Wang, T.Q., Yu, Q, Li, Y.M., Shi,
W., Tip clearance flow in a transonic compressor rotor,
Journal of Engineering Thermophysics, Vol. 21, No. 5,
2000, pp. 570572.

You might also like