You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/222155917

Counter-rotating streamwise vortex formation in the turbine cascade


with endwall fence

Article in Computers & Fluids · May 2001


DOI: 10.1016/S0045-7930(00)00026-8

CITATIONS READS
39 1,241

2 authors:

Young Moon Seong-Ryong Koh


Korea University Forschungszentrum Jülich
82 PUBLICATIONS 1,665 CITATIONS 59 PUBLICATIONS 494 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Young Moon on 31 December 2018.

The user has requested enhancement of the downloaded file.


Computers & Fluids 30 (2001) 473±490
www.elsevier.com/locate/comp¯uid

Counter-rotating streamwise vortex formation in the turbine


cascade with endwall fence
Young J. Moon *, Sung-Ryong Koh
Department of Mechanical Engineering, Korea University, 1-5 Anam-dong, Sungbuk-ku, Seoul, 136-701, South Korea
Received 29 March 1999; received in revised form 27 January 2000; accepted 6 June 2000

Abstract
The three-dimensional turbulent ¯ows in the turbine cascade with and without endwall fences are nu-
merically investigated by solving the incompressible Navier±Stokes equations with a high-Reynolds
number k± turbulence closure model. The limiting streamline patterns and the static pressure contours at
the suction surface of the blade as well as on the cascade endwall are employed to visualize the e€ectiveness
of the endwall fence for the secondary ¯ow control. The analysis on the streamwise vorticity contour maps
along the cascade with the three-dimensional representation of their iso-surfaces reveals the complicated
structures of the vortical ¯ows in the turbine cascade with endwall fence, and also leads to an understanding
on the formation of a counter-rotating streamwise vortex over the fence. The mechanisms of controlling the
secondary ¯ow and the proper selection of an optimal fence height are also explained. Ó 2001 Elsevier
Science Ltd. All rights reserved.

Keywords: Boundary layer fence; Streamwise vortex; Secondary ¯ow; Turbine cascade

1. Introduction

In the three-dimensional turbine cascade, a secondary ¯ow is generated by the interaction of a


horseshoe vortex upstream from the leading edge of the blade with a passage vortex developed by
the curvature e€ect of the bend. The pressure-side leg of the horseshoe vortex rotates in the same
direction as the passage vortex, and as a result these two are superposed as schematically shown in
Fig. 1. This secondary ¯ow development a€ects the aerodynamic performance of the turbine, and
also increases the heat transfer rate from the hot ¯uid to the blade and the endwall surfaces. A
comprehensive review on the secondary ¯ow structures in the turbine cascade has been presented

*
Corresponding author. Tel.: +82-2-3290-3358; fax: +82-2-926-9290.
E-mail address: yjmoon@korea.ac.kr (Y.J. Moon).

0045-7930/01/$ - see front matter Ó 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 4 5 - 7 9 3 0 ( 0 0 ) 0 0 0 2 6 - 8
474 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Fig. 1. Vortex structure in the turbine cascade.

by Sieverding [1], and various experimental studies have also been conducted by Jabbari [2],
Langston [3], Yamamoto [4,5], Sharma [6], and Goldstein [7].
For control of the secondary ¯ow, Kawai [8±11], Chung [12], and Lee [13] have experi-
mentally investigated on a use of the boundary layer fence at the endwall to hinder the merger
of the horseshoe vortex with the passage vortex, and have measured the total pressure losses
and velocity distributions in the wake region. An optimal size and location of the boundary
layer fence were determined by observing the fact that a fence with 1/3 height of the inlet
boundary layer thickness posted in the middle between blades reduces the secondary ¯ow losses
by 20%.
In the present study, the incompressible Navier±Stokes equations are numerically solved with
a high Reynolds number k± turbulence closure model for investigating the three-dimensional
turbulent ¯ows of turbine cascade with and without the endwall fences. Our concerns lie on the
accurate numerical prediction of the secondary ¯ow in the cascade as well as on the under-
standing of its complex three-dimensional ¯ow structures. Main objectives are to investigate how
the boundary layer fence reduces the secondary ¯ow development in the cascade, and to un-
derstand why the fence with 1/3 height of the inlet boundary layer thickness is optimal for the
secondary ¯ow control. The limiting streamline patterns and the static pressure contours at the
suction surface of the blade and the endwall are employed to visualize the topology of the three-
dimensional secondary ¯ow structures in the turbine cascade with and without endwall fences.
The streamwise vorticity contours at the various stations in the passage are analyzed to un-
derstand how the pressure-side leg of the horseshoe vortex interacts with the endwall fence of
di€erent heights and its e€ects on the development of the passage vortex near the suction-side
blade.
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 475

2. Computational methods

2.1. Governing equations

A non-dimensional conservative form of the three-dimensional incompressible Navier±Stokes


equations is expressed in the generalized coordinates as
     
o U o V o W
‡ ‡ ˆ 0; 1†
on J og J of J

^ oE^ oF^ oG
oQ ^ oE^v oF^v oG ^v
‡ ‡ ‡ ˆ ‡ ‡ ; 2†
ot on og of on og of
where Q^ ˆ 1=J ‰u; v; wŠT and
2 3 2 3 2 3
uU ‡ nx p uV ‡ gx p uW ‡ fx p
1 1 ^ ˆ 1 4 vW ‡ fy p 5:
E^ ˆ 4 vU ‡ ny p 5; F^ ˆ 4 vV ‡ gy p 5; G 3†
J J J
wU ‡ nz p wV ‡ gz p wW ‡ fz p
^ represents the dependent variable vector for the momentum equations, E,
Here, Q ^ F^, and G^ the
convective ¯ux vectors, U, V, and W the contravariant velocity components, and J, a Jacobian of
the metric transformation.
^v can be written in a tensor form,
The viscous ¯ux vectors E^v , F^v , and G
  
meff on oui onl ouj onl
E vi ˆ ‡ ;
Re J oxj onl oxj onl oxi
  
meff og oui onl ouj onl
F vi ˆ ‡ ; 4†
Re J oxj onl oxj onl oxi
  
meff of oui onl ouj onl
G vi ˆ ‡ ;
Re J oxj onl oxj onl oxi

where the e€ective viscosity is de®ned as meff ˆ ml ‡ mt , and ml and mt correspond to the laminar and
turbulent viscosities, respectively.
A turbulence closure is made by solving the standard k± turbulence transport equations. A
non-dimensional form of the k± equations is also written in the generalized coordinates as

oQ^ t oE^t oF^t oG^t oE^vt oF^vt oG^vt


‡ ‡ ‡ ˆ ‡ ‡ ‡ S^t ; 5†
ot on og of on og of
where the turbulence transport variables and the convective and viscous ¯ux vectors are de®ned as
       
^t ˆ 1 k 1 kU 1 kV ^t ˆ 1 kW
Q ; E^t ˆ ; F^t ˆ ; G ; 6†
J e J eU J eV J eW

and
476 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490
2  n o3
mt
1 6 rn  rn† ok ok
‡ rn  rg† og ‡ rn  rf† ok
rk on of 7
E^vt ˆ 4  n o5
Re J mt oe
rn  rn† on ‡ oe
rn  rg† og oe
‡ rn  rf† of
re
2  n o3
mt
rg  rn† ok ‡ ok
rg  rg† og ‡ rg  rf† ok
1 6 rk on of 7
F^vt ˆ 4  n o 5: 7†
Re J mt oe
rg  rn† on ‡ rg  oe
rg† og ‡ rg  oe
rf† of
re
2  n o3
mt
1 rf  rn† ok ‡ rf  rg† ok ‡ rf  rf† ok
^vt ˆ 6 rk on og of 7
G 4  n o5
Re J mt oe
rf  rn† on ‡ oe
rf  rg† og oe
‡ rf  rf† of
re

The source vector is given by


 
^ 1 G e
St ˆ C1 e 2 ; 8†
J k
G C2ke
where the generation term G is
 
oui onl ouj onl oui onl
G ˆ mt ‡ ; 9†
onl oxj onl oxi onl oxj
and the turbulent viscosity is determined by
 
Cl k 2
mt ˆ min ; ml : 10†
e
The standard constants of
Cl ˆ 0:09; C1 ˆ 1:44; C2 ˆ 1:92; rk ˆ 1; re ˆ 1:217 11†

are used and a wall function is employed for the wall treatment.

2.2. Numerical algorithms

The incompressible Navier±Stokes equations are solved in a coupled manner by a projection


method based procedure [14]. Even though this method was primarily used for unsteady ¯ow
computations [15,16], exploits to the steady ¯ow applications have been sought by Sotiropoulos
et al. [17,18], employing the numerical methods [19±21] widely used in the time-marching solution
procedure; for example, a multi-stage Runge±Kutta method for time integration of the mo-
mentum equations and the implicit residual smoothing and local time stepping techniques for the
convergence acceleration. In this study, a similar numerical approach is used for calculating the
three-dimensional turbulent ¯ows in the turbine cascade, and some numerical observations made
on the properties of the present method are discussed.
The solution procedures are brie¯y described here. At a given time step n,

(i) calculate the intermediate velocity Q from the momentum equations Eq. (2) in an explicit
manner
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 477

Q Qn
ˆ f Qn ; pn †; 12†
Dt
(ii) solve the Poisson equation for a potential function w ˆ Dt dp†
r2 w ˆ r  Q ; 13†
n‡1
ensuring the continuity equation r  Q ˆ 0,
(iii) update the pressure from
w
pn‡1 ˆ pn ‡ ; 14†
Dt
(iv) the velocity at n ‡ 1 time step is obtained by
Qn‡1 ˆ Q rw: 15†

A spatial discretization on the governing equations is then made by following the cell-center
based ®nite-volume method with a non-staggered arrangement of the variables. A conservative
form of the momentum equations Eq. (2) is cast into a discrete form as
oQ h i
ˆ J dn E^ E^v † ‡ dg F^ F^v † ‡ df G ^ G ^v † ; 16†
ot i;j;k

ˆ R Q†i;j;k ;
where, for example, dn † ˆ †i‡1=2;j;k †i 1=2;j;k , and the right-hand side term R Q†i;j;k repre-
sents the residual vector.
Here, a MUSCL approach [22] is used to apply the second-order upwind-di€erencing scheme
to the convective ¯ux terms in Eq. (16). For example, a numerical ¯ux at the cell face in the
n-direction is determined by
E^i‡1=2 ˆ E^ Q
i‡1=2 ; U i‡1=2 ; p i‡1=2 †; 17†
where
Qi Qi 1

Qi‡1=2 ˆ Qi ‡ U P 0†;
2
Qi‡2 Qi‡1
 i‡1=2 18†

i‡1=2 ˆ Qi‡1 2
U i‡1=2 < 0†:
The transported variables Q i‡1=2 at the cell face are determined by a linear extrapolation from
each upwind direction based on a sign of the contravariant material velocity at the cell face. The
contravariant material velocity and pressure at the cell face are, however, de®ned by a linear
interpolation between two neighboring cells. This procedure may lead to an odd±even decoupling
of the pressure ®eld and therefore a momentum linear interpolation technique [23] was applied
after solving Eq. (12). The convective terms in the other directions as well as in the k± turbulence
equations are upwind-di€erenced in the same manner, while the terms related to the di€usion
process are centrally di€erenced.
For the time integration of the momentum equations, Sotiropoulos et al. [17,18] employed the
numerical techniques widely used in the time-marching solution methods for the compressible
Navier±Stokes equations [19] or in the arti®cial compressibility method [24] for the incompressible
Navier±Stokes equations [20,21]; for example, a multi-stage Runge±Kutta method with implicit
residual smoothing and local time stepping techniques. The computational eciency largely
478 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

obtained by the multi-stage Runge±Kutta method in most of the external aerodynamics calcu-
lations was, however, not fully achieved as expected to be, especially when calculating the internal
turbulent ¯ows of turbomachinery coupled with the k± turbulence equations. In this regard, a
forward Euler method is employed for the time advancement of the momentum equations Eq.
(16), and is subject to be further discussed on its convergence behavior in Section 2.3.
For the convergence acceleration to a steady state, the implicit residual smoothing and the local
time stepping are used in conjunction with the time-integrator. The original residual R Q†i;j;k in
Eq. (16) is ®ltered at each iteration by solving a di€usion-like equation,

1 e dnn † 1 e dgg 1 e dff †Ri;j;k ˆ Ri;j;k ; 19†
where dnn , dgg , dff are the standard second-di€erence operators. For a smoothing coecient ,
values between 0.5 and 1.5 are used in practice.
Also, a local time step Dti;j;k is determined by a harmonic average of the time steps in each
direction,
1 1 1 1
ˆ ‡ ‡ ; 20†
Dti;j;k Dtn Dtg Dtf i;j;k

and a local time step in the n-direction, for example, is de®ned by

k
D tn ˆ q ; 21†
jU j ‡ b n2x ‡ n2y ‡ n2z
i;j;k
p
where k is a CFL number and b ˆ 1 ‡ u2 ‡ v2 ‡ w2 denotes an arti®cial speed of sound.
The locally de®ned time steps substantially accelerates the convergence rate not only by al-
lowing spatially di€erent time advancements but also by the enhancement of the numerical sta-
bility. The latter is due to the fact that the pressure ®eld is updated by Eq. (14), in which the
potential function w and the locally de®ned time step Dti;j;k are both proportional to the mesh size,
yielding w=Dti;j;k stable. The extended numerical stability of using the implicit residual smoothing
and the local time stepping with the present segregated solution procedure will be further dis-
cussed with a demonstrated result in Section 2.3.
In the present study, the Poisson equation Eq. (13) is implicitly solved by an alternate direc-
tional implicit (ADI) method with a single sweep, and a Neumann type boundary condition is
used for w. The turbulence transport equations (5)±(11) are also solved by following the same
numerical procedures described in this section.

2.3. Computations and validation

A turbine cascade in the experiment of Wang et al. [25] is considered for a validation study. The
blade pro®le and its coordinate system are illustrated in Fig. 2, where x and y indicate the axial
and pitch directions and z the spanwise direction. The blade de®nition is given in Ref. [25], and its
geometrical details are summarized in Table 1.
A three-dimensional turbulent cascade ¯ow is computed for Reex ˆ 5:4  105 , and the sub-
scription ex denotes a basis for an exit velocity. Fig. 3 shows a perspective view of the computa-
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 479

Fig. 2. Turbine blade de®nitions and coordinate system.

Table 1
Turbine cascade geometry
Chord length (C) 184.15 mm
Axial chord to chord ratio (Cax =C ) 0.704
Aspect ratio (H =C ) 2.483
Solidity (C=P ) 0.75
Blade inlet angle (b1 ) 35°
Blade outlet angle (b2 ) 72:5°
Turning angle 107:5°
Incidence angle 0°

tional grids within the blade passage and the mesh details near the blade leading and trailing edges.
The incoming ¯ow is turbulent with a boundary layer thickness of 24 mm and a turbulent intensity
of 0.7%. A standard wall-function is used on the rigid walls when the k± turbulence equations are
solved. At the out¯ow plane, the streamwise gradients of all variables are set to zero with a velocity
correction for the global mass conservation. The mid-span plane is treated as a symmetric plane
and a periodic boundary condition is applied at the upstream and downstream side planes.
480 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Fig. 3. Computational grids of a turbine cascade. (a) Surface grids of a blade passage 75  30  25† and (b) mesh
details near the leading and trailing edges.

A typical convergence behavior of the present numerical method is shown in Fig. 4 for the
computation of a three-dimensional turbulent cascade ¯ow. Here, a CFL number of 1 was used
with the implicit residual smoothing and local time stepping techniques, and a six-order of the
magnitude in residual drops in 4000 iterations. It is interesting to note that a time step allowed by
the forward Euler method in the fully coupled formulations (discussed in Section 2.2) is usually
much smaller than the one used here. It seems that the numerical stability using the local time
stepping in the segregated solution procedure seems far less constricted, whereas, to that governed
by a fully coupled system of equations for the pressure and the velocities.
An iso-surface of the non-dimensional streamwise vorticity (Cxs =U ˆ 5) is shown in Fig. 5,
which represents well the three-dimensional nature of the turbine cascade ¯ow. This ®gure clearly
depicts a secondary ¯ow structure within the cascade, i.e. the merger of the pressure-side leg of a
horseshoe vortex generated upstream of the blade leading edge with a passage vortex developed
within a curved cascade blade. The ¯ow patterns and the size of the secondary ¯ow are also
distinctively identi®able in Fig. 9 from the limiting streamlines on the cascade endwall and the
suction surface of the blade.
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 481

Fig. 4. Convergence of a turbine cascade computation.

Fig. 5. An iso-surface of the non-dimensional streamwise vorticity (Cxs =U ˆ 5), visualizing the structure of the sec-
ondary ¯ow.
482 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Accuracy assessments of the present method are also made by comparing the computed so-
lutions with some experimental data available. First, the computed static pressure distributions at
the blade mid-span are compared with the data measured by Wang et al. [25]. Fig. 6 shows results
of the grid dependence test conducted with various grid systems. The solutions obtained by using
a two-dimensional approximation are also closely compared with measurement, but not quite
accurately as the three-dimensional ones. This discrepancy protrudes at the suction surface of the
cascade near the trailing edge, due to the e€ect of the secondary ¯ow development. The three-
dimensional grid of 95  40  40 points in axial, pitch, and spanwise directions yields the closest
solution to the experiment. Second, the total pressure loss contours at a cross-section near the
trailing edge are compared with experimental data measured by Lee [13]. In this experiment, the
same blade pro®le and ¯ow conditions as Wang et al.'s were used except that an aspect ratio of
the blade was modi®ed to 2.0. Fig. 7 shows a comparison between computation and experiment
for the two cases of turbine, with and without the endwall fences, and Fig. 7(b) corresponds to the
case of an the endwall fence (hf ˆ 1=3†d). The computed results of total pressure loss contours
agree reasonably well with experiment, in terms of the size and location of the secondary ¯ow
developed near the suction surface of the blade.

3. Secondary ¯ow structures in the turbine cascade

Understandings on how the boundary layer fence interacts with the right-hand leg of the
horseshoe vortex and as a result reduces the secondary ¯ow development in the cascade are

Fig. 6. Comparison of static pressure coecients (grid dependence test).


Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 483

(a)

(b)

Fig. 7. Comparison of total pressure loss coecient contours. (a: endwall; b: midspan). (a) No fence and (b) with a
fence (hf ˆ 1=3†d).

pursued in the present numerical study. A fence of 5 mm thickness is mounted on the cascade
endwall to suppress the secondary ¯ow development by hindering the interaction of an upcoming
484 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Fig. 8. An endwall fence (hf ˆ 3=3†d) modeled on a computational grid.

horseshoe vortex from the turbine leading edge with the passage vortex. The boundary layer fence
pro®le is same as the blade camberline, and three fences of 1/3, 2/3, and 3/3 heights of the inlet
boundary layer thickness are tested as in the experiments [8±13]. Fig. 8 shows an endwall fence of
the highest case (hf ˆ 3=3†d) modeled on a computational grid, and hf denotes the fence height
and d the inlet boundary layer thickness. Same ¯ow conditions are also used as in the previous
calculations (described in Section 2.3), and computations are made for the blade with an aspect
ratio of 2.
The static pressure coecient contours and limiting streamlines presented in Fig. 9 reveal the
structures of the three-dimensional cascade ¯ows with and without endwall fences. From Fig. 9(a)
and (b), one can clearly recognize the size of the secondary ¯ow developed within the cascade for
each case. With an endwall fence the secondary ¯ow development near the suction surface is much
suppressed, and an optimal fence height for the secondary ¯ow control is clearly discernable from
these ®gures. From the limiting streamline patterns on the endwall shown in Fig. 9(c), one can see
the ¯ow topology near the endwall surface and understand how the ¯ow structure is altered by
inserting an endwall fence. A saddle point is clearly visible near the blade leading edge due to the
horseshoe vortex formation, and the right-hand leg of the horseshoe vortex leaning to the suction
side is blocked by the fence. It can also be noticed that a strong cross-¯ow from the pressure side
to the suction side near the cascade endwall is considerably reduced by the boundary layer fence,
and that, as the fence height increases, the cross-¯ow becomes stronger at the region between the
fence and the suction surface. With the fence of 1/3 height of the inlet boundary layer thickness,
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 485

Fig. 9. Visualizations of the secondary ¯ow structures in turbine cascades with and without endwall fences. (a) Contour
plots of static pressure coecient at the suction surface (a: endwall, b: mid-span), (b) limiting streamlines at the suction
surface and (c) limiting streamlines on the endwall.

the secondary ¯ow is least developed, which, in fact, agrees well with the experimental observa-
tions made by Kawai [8±11], Chung [12], and Lee [13].
One can also observe that, as the fence height increases, the blocking point of the divided
streamline moves downstream due to the blockage e€ect. It is known that the secondary ¯ow
development is mainly due to the merger of the horseshoe vortex with the passage vortex.
486 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Therefore, it may be simply conjectured that a complete blocking of the horseshoe vortex will
reduce the secondary ¯ow development, since the merger of these two counter-clockwise rotating
vortexes is prohibited. On the contrary, the secondary ¯ow near the suction surface becomes
stronger, as the fence height increases. This opposite phenomenon may be understood from Fig.
10 with the non-dimensional streamwise vorticity maps at the stations of 60%, 70%, 80%, and 90%
of the axial chord length along the cascade. Fig. 10(a) clearly shows the formation of a strongly
merged streamwise vortex developed in the downstream of the cascade with no endwall fence. In
fact, it is important to notice that the size of the merged vortex at the 60% location of the axial
chord length is in a good proximity to the inlet boundary layer thickness of 24 mm. Therefore, it is
useful to use the inlet boundary layer thickness as a characteristic length scale, and as a result the
fence height relative to this thickness plays an important role in control of the streamwise vortex
formation. From Figs. 10(b)±(d), it may be understood why the fence with 1/3 height of the inlet
boundary layer thickness generates the least streamwise vorticity, comparing with the other two
cases. Since the size of the merged vortex is in the order of an inlet boundary layer thickness, the
horseshoe vortex is almost blocked in the region between the pressure side and the fence for
hf ˆ 3=3†d, and therefore, a stand-alone passage vortex revives at the suction side. This is called a
``half-pitch e€ect''. On the other hand, if the fence height is lower than the boundary layer
thickness, the counter-clockwise rotating (in solid lines) horseshoe vortex tops over the fence and
produces a clockwise rotating streamwise vortex (in broken lines) at the region between the fence
and the suction surface. The rotational direction of this newly generated streamwise vortex is, in
fact, opposite to that of the passage vortex, and therefore it plays an important role in reducing
the development of the secondary ¯ow near the suction surface of the blade. As one can see from
Fig. 10(b), the fence with 1/3 height of the inlet boundary layer thickness most e€ectively generates
the counter-rotating streamwise vortex, and therefore the development of the passage vortex near
the suction surface is most suppressed. The interaction mechanisms of the horseshoe vortex
with the boundary layer fence and the formation of a counter-rotating streamwise vortex in the
cascade are also three-dimensionally visualized in Fig. 11 by the iso-surfaces of the non-dimen-
sional streamwise vorticity of ‡15 (counter-clockwise rotating horseshoe and passage vortexes
in dark color) and 15 (clockwise rotating streamwise vortex in light color). The large scale
structure of the merged passage vortex rotating in the positive direction is clearly shown in
Fig. 11(a) for the cascade with no endwall fence, whereas, for the case of hf ˆ 1=3†d shown in
Fig. 11(b), one can see the right-hand leg of the horseshoe vortex continuously scraping the fence
edge from the upstream and generating a counter-rotating streamwise vortex at the region be-
tween the fence and the suction surface. Fig. 11(d) shows that, for the case with the highest fence,
the horseshoe vortex is almost completely blocked by the fence and there is no formation of a
counter-rotating streamwise vortex in the cascade. The e€ectiveness of reducing the passage
vortex near the suction surface (shown as a dark iso-surface of the positive streamwise vorticity) is
clearly noticeable among the cases from Fig. 11(a)±(d).

4. Concluding remarks

The three-dimensional turbulent cascade ¯ows with and without endwall fences are numerically
investigated by solving the incompressible Navier±Stokes equations with a high Reynolds number
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 487

Fig. 10. Comparison of the non-dimensional streamwise vorticity contours along the cascade. (a: endwall, b: 1/3 mid-
span, contour level increment ˆ 5). (a) No fence, (b) hf ˆ 1=3†d fence, (c) hf ˆ 2=3†d fence and (d) hf ˆ 3=3†d fence.

k± turbulence closure model. The numerical accuracy of the present method was validated by
comparing the static pressure distributions at the mid-span of the cascade and the total pressure
loss contours at a cross-section near the trailing edge with experimentally measured data, and also
by the fact that the computed results strongly support the experimental observations that an
optimal fence height for the least secondary ¯ow is one-third of the inlet boundary layer thickness.
The three-dimensional nature of the secondary ¯ow in the turbine cascade is also understood by
the visualizations of the limiting streamline patterns and the static pressure contours at the suction
surface of the blade as well as on the cascade endwall. Analysis on the non-dimensional
streamwise vorticity contour maps along the cascade with the three-dimensional representation of
488 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

Fig. 11. Iso-surfaces of the non-dimensional streamwise vorticity (Cxs =U ˆ 15; 15) along the cascade ¯ow passage,
illustrating the formation of a counter-rotating streamwise vortex. (a) No fence, (b) hf ˆ 1=3†d fence, (c) hf ˆ 2=3†d
fence and (d) hf ˆ 2=3†d fence.

their iso-surfaces reveals the structure of the complicated vortical ¯ows in the turbine cascade with
endwall fence, and explains the following:

1. selection of the inlet boundary layer thickness as a characteristic length scale for control of the
secondary ¯ow in turbine cascade,
2. mechanisms of reducing the passage vortex development near the suction surface by controlling
the fence height relative to the inlet boundary layer thickness,
3. the role of the boundary layer fence on formation of the counter-rotating streamwise vortex,
and
Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490 489

4. its e€ectiveness of reducing the secondary ¯ow by down-sizing the fence height to 1/3 of the
inlet boundary layer thickness.

It leads to a conclusion that understandings on formation of the counter-rotating streamwise


vortex in the turbine cascade with endwall fence explain the mechanisms of controlling the sec-
ondary ¯ow development, and also for a proper selection of an optimal fence height.

Acknowledgements

Authors would like to express gratitude to Mr. Lee, Yong-Jin and Prof. Chung, Jin-Taek at
Korea University for providing experimental data, which are used for assessments of the present
computational results.

References

[1] Sieverding CH. Recent progress in the understanding of basic aspects of secondary ¯ows in turbine blade passages.
J Engng Gas Turbines Power 1985;107:248±57.
[2] Jabbari MY, Goldstein RJM, Marston KC, Eckert ERG. Three dimensional ¯ow within large scale turbine
cascades. Warme und Sto€ ubertragung 1992;27:51±9.
[3] Langston LS, Nice ML, Hooper RH. Three dimensional ¯ow within a turbine cascade passage. J Engng Power
1977;99:21±8.
[4] Yamamoto A. Production and development of secondary ¯ows and losses in two types of straight turbine cascades:
Part 1 ± a stator case. J Turbomach 1987;109:186±93.
[5] Yamamoto A. Production and development of secondary ¯ows and losses in two types of straight turbine cascades:
Part 2 ± a rotor case. J Turbomach 1987;109:194±200.
[6] Sharma OP, Butler TL. Predictions of endwall losses and secondary ¯ows in axial ¯ow turbine cascades.
J Turbomach 1987;109:229±36.
[7] Goldstein RJ, Spores RA. Turbulent transport on the endwall in the region between adjacent turbine cascades.
J Heat Trans 1988;110:862±9.
[8] Kawai T, Adachi T, Shinoki S. Secondary ¯ow control and loss reduction in a turbine cascade using endwall fences.
JSME Int J Ser II 1989;32(3):375±87.
[9] Kawai T, Adachi T, Shinoki S. Visualization study of three-dimensional ¯ows in a turbine cascade endwall region.
JSME Int J Ser II 1989;33(2):256±64.
[10] Kawai T, Adachi T, Shinoki S. Improvement in turbine blade aerodynamic force in the tip region. JSME Int J Ser
II 1990;33(3):517±24.
[11] Kawai T. E€ect of combined boundary layer fences on turbine secondary ¯ow and losses. JSME Int J Ser B
1994;37(2):377±84.
[12] Chung JT, Simon TW, Buddhavarapu T. Three-dimensional ¯ow near the blade/endwall junction of a gas turbine:
application of a boundary layer fence. ASME Paper 91±GT±45, 1991.
[13] Lee YJ. Experimental study on e€ects of the boundary layer fence on the three-dimensional ¯ow in gas turbines.
MS Thesis, Department of Mechanical Engineering, Korea University, 1997.
[14] Chorin AJ. Numerical solution of the Navier±Stokes equations. Math Comput 1968;22:745±62.
[15] Hallow FH, Welch JE. Numerical calculation of time dependent viscous incompressible ¯ow of ¯uid with free
surface. Phys Fluids 1965;8:2182±9.
[16] Hirt CW, Cook JL. Calculating three-dimensional ¯ows around structures and over rough terrain. J Comput Phys
1972;10:324±40.
490 Y.J. Moon, S.-R. Koh / Computers & Fluids 30 (2001) 473±490

[17] Sotiropoulos F, Abdallah S. A primitive variable method for the solution of three-dimensional incompressible
viscous ¯ows. J Comput Phys 1992;103:336±49.
[18] Sotiropoulos F, Kim WJ, Patel VC. A computational comparison of two incompressible Navier±Stokes solvers in
three-dimensional laminar ¯ows. Comput Fluids 1994;23(4):627±46.
[19] Jameson A, Schmidt W, Turkel E. Numerical solutions of the Euler equations by ®nite volume methods using
Runge±Kutta time-stepping schemes. AIAA Paper No. 81-1259, 1981.
[20] Farmer J, Martinelli L, Jameson A. Fast multigrid method for solving incompressible hydrodynamic problems
with free surfaces. AIAA J 1994;32(11):1175±82.
[21] Moon YJ. Explicit and unstructured grid solution of the incompressible Navier±Stokes equations using arti®cial
compressibility method. Comput Fluid Dynam J 1997;6(1):39±50.
[22] Van Leer B. Towards the ultimate conservative di€erence scheme, V. second-order sequel to Godunov's method.
J Comput Phys 1979;32:101±36.
[23] Rhie CM, Chow WL. Numerical study of the turbulent ¯ow past an airfoil with trailing edge separation. AIAA J
1983;21(11):1525±32.
[24] Chorin AJ. A numerical method for solving incompressible viscous ¯ow problems. J Comput Phys 1967;2:12±26.
[25] Wang H, Olson SJ, Goldstein RJ, Eckert ERG. Flow visualization in a linear turbine cascade of high performance
turbine blades. ASME Paper 95±GT±7, 1995.

View publication stats

You might also like