Numerical Simulation of Turbine Blade B.L & Heat Transfer & Assesment of Turbulence Models

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Numerical Simulation of Turbine

Blade Boundary Layer and Heat


J. Luo1 Transfer and Assessment of
Graduate Student.
Mem. ASME Turbulence Models
B. Lakshminarayana The boundary layer development and convective heat transfer on transonic turbine
nozzle vanes are investigated using a compressible Navier-Stokes code with three
Evan Pugh Professor and Director.
low-Reynolds-number k- e models. The mean-flow and turbulence transport equations
Fellow ASME
are integrated by a four-stage Runge-Kutta scheme. Numerical predictions are com-
pared with the experimental data acquired at Allison Engine Company. An assessment
Center for Gas Turbine and Power, of the performance of various turbulence models is carried out. The two modes of
The Pennsylvania State University,
University Park, PA 16802
transition, bypass transition and separation-induced transition, are studied compara-
tively. Effects of blade surface pressure gradients, free-stream turbulence level, and
Reynolds number on the blade boundary layer development, particularly transition
onset, are examined. Predictions from a parabolic boundary layer code are included
for comparison with those from the elliptic Navier-Stokes code. The present study
indicates that the turbine external heat transfer, under real engine conditions, can
be predicted well by the Navier-Stokes procedure with the low-Reynolds-number
k- e models employed.

Introduction sensitive to the starting profile and conditions specified at the


edge of the boundary layer. The prediction of transition by a
An improved understanding of and predictive capability for boundary layer code (with LRN k- e models) is particularly
turbine viscous flowfield and heat transfer is crucial for achiev- sensitive to the initial profile of e, which is not well defined
ing higher turbine inlet temperatures and for minimizing various and not available from experimental measurements, thus in-
aerodynamic losses. The turbine flow and thermal fields are creasing the uncertainty about the prediction accuracy (Savill,
dominated by complex flow phenomena including laminar, tran- 1993). For separated flows or three-dimensional endwall flows,
sitional, and turbulent flows; stagnation flow; large pressure Navier-Stokes procedures must be used. There have been some
gradients and free-stream turbulence; three-dimensional vortical attempts using Navier-Stokes procedures with LRN k- e mod-
flows; and unsteady separated flow at the trailing edge. Simo- els to compute turbine heat transfer, e.g., Hah (1989), Ameri
neau and Simon (1993) reviewed the progress in predicting and Arnone (1994), and Luo and Lakshminarayana (1995).
heat transfer in the turbine gas path in three major areas: transi- However, mixed success has been achieved; there is no consis-
tion (for midspan flows), three-dimensional endwall flows, and tency in the quality of predictions for various test cases used.
film-cooling. A comprehensive review of transition in gas tur- The objective of this paper is to develop a numerical proce-
bines was presented by Mayle (1991), in which three modes of dure for computation of coupled turbine aerodynamics and heat
transition: natural transition, bypass transition, and separation- transfer, and to assess the accuracy and suitability of various
induced transition were discussed. For transition at high free- turbulence models for predicting transition and heat transfer.
stream turbulence levels, such as is typically found in turbines, Two modes of transition, the bypass transition and the separa-
the linear stability mechanism (i.e., the two-dimensional Toll- tion-induced transition, are studied comparatively. Numerical
mien-Schlichting instability wave) is bypassed such that the predictions are compared with the experimental data obtained
turbulence spots are directly produced, which leads to the so- by Hylton et al. (1983) at Allison Engine Company. A two-
called bypass transition. The third transition mode is caused by dimensional boundary layer code, TEXSTAN (Crawford,
the separation of laminar boundary layer, which may occur in 1985), was employed to predict heat transfer for comparison
flows subjected to adverse pressure gradients. with the Navier-Stokes solutions.
Schmidt and Patankar (1992) presented a detailed analysis
of the mechanism by which low-Reynolds-number (LRN) k-
e models simulate the transition process. Savill (1993) coordi- Turbulence Models and Numerical Techniques
nated a comprehensive comparison of several test cases of flat- The compressible Navier-Stokes equations in conservative
plate boundary layers undergoing bypass transition with various form are utilized, with the effective stress tensor r„ and the
LRN turbulence models. A number of researchers have applied effective heat flux vector qt given by
LRN k- e models to calculate convective heat transfer to turbine
blading, e.g., Hylton et al. (1983), and Schmidt and Patanker Tlij pUiUj = (fj,i + /!,)[(dUi/dXj + dUj/dXi)
(1992). These computations are performed with boundary-layer
codes, which are efficient and can provide high resolution of j6vdUk/dxk. I 6upk (1)
the viscous layer. However, the boundary layer solutions are
dT
d (2)
'\Pr, PrJ dx,
1
Current address: Solar Turbines Inc., CA.
Contributed by the Heat Transfer Committee for publication in the JOURNAL
OF TURBOM ACHINERY . Manuscript received by the Heat Transfer Committee Octo- where p., is the turbulent eddy viscosity, which is computed
ber 1, 1996. Associate Technical Editor: M. O. Dunn. using the k— e model; Pr, = 0.9.

794 / Vol. 119, OCTOBER 1997 Transactions of the ASME

Copyright © 1997 by ASME


Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo
Table 1 Low-Reynolds-number functions used in CH, LB, and FLB yana, and Barnett (1993, denoted as FLB). Chien's k-e model
models
is popular in Navier-Stokes solvers due to its good numerical
stability, which can be attributed to its near-wall asymptotic
Model f^ f, f2 consistency, as shown by Patel et al. (1985) and the algebraic
CH I-exp(-O.OI15y+) 1.0 l-0.22exp(-ReJ736)) form of its near-wall terms. However, the important low-Reyn-
LB [l-cxp(-0.0165Rey)]2(l+20.5/Ret) l+(0.O6/t'M)3 l-exp(-Rc2) olds-number function /^ in this model depends only on _y+,
FLB aMJyjKe,-* 1.0 (l-0.22exp(-Re2/36))f2
which makes it inaccurate for separated flows. The LB model
( 1 - 0 . 4 ^ Ret) [ 1 - e x p l - R e ^ ^ ) ] 3 does not contain y * in the low-Reynolds number functions, but
it is not asymptotically consistent in the near-wall region (Patel
et a l , 1985). Fan et al. (1993) developed an improved LRN
k- e model with near-wall asymptotic consistency. This model
Low-Reynolds-Number k-e Model. In the low-Reyn- has been shown to capture wake-induced transition on compres-
olds-number k-e model, the eddy viscosity is obtained from sor blades quite accurately (Fan and Lakshminarayana, 1996).
the Prandtl-Kolmogorov relation: Table 1 gives the near-wall functions in the three models
described above, where the near-wall function in the FLB model
is given by
M< = CJ„p '• (3)
y VRev
where C^ = 0.09 and fM is a near-wall damping function given L= \ - exp-^ - +
in Table 1. The low-Reynolds-number k-e equations can be
1
2.30 2.30 8.89
written as
Rey
d(pk) d(PU,k) _ d 1 — exp (6)
20
dt dxj dxj _\ akJ dXj_
Numerical Procedure
+ ppt - p(e + D) (4)
Navier-Stokes Code. The Navier-Stokes code (hereafter
d(pe) | d(pU,e) d
referred to as RK2D, Kunz and Lakshminarayana, 1992) uses
dt dxj Ox, \ a, J dxj _ a standard four-stage Runge-Kutta scheme. Fourth-order arti-
ficial dissipation is included to damp high wave number errors
and second-order artificial dissipation is added to improve the
+ p-[C£lflPk -Cafie] + pE (5) shock capturing. Second-order artificial dissipation is also added
k
to the k and e equations to prevent odd-even decoupling since
where D, £ , / , and/ 2 are near-wall terms. the convection terms in Eqs. (4) and (5) are discretized with
Three representative LRN k- e models are evaluated: Chien's second-order accurate central differences. Multiplied by the
(1982) version (denoted CH), Lam-Bremhorst's (1981) ver- square of the normalized local velocity, (U/U„)2, the smooth-
sion (denoted LB); and a recent model from Fan, Lakshminara- ing in both mean-flow and turbulence equations was reduced

Nomenclature
ARC = distance along airfoil sur- L = turbulence length scale utUj = Reynolds stress
face from stagnation point M = Mach number x = coordinate along axial chord
to trailing edge P = pressure y coordinate in tangential direc-
Cf = skin friction coefficient Pk = production rate of turbulent tion (pitchwise)/normal dis-
(nondimensionalized by kinetic energy = -«,«, (dUJ tance to the wall
local free-stream dynamic dxj) = wall distance variable = yuTlu
head) Pr,, Pr, = laminar and turbulent Prandtl r
= Kronecker delta
Cp, Cv = specific heat at constant numbers = boundary layer and displace-
pressure and volume PS = pressure surface , 5 * ment thicknesses
C/j, C\, C2, q> = Cartesian components of heat = dissipation rate of turbulent ki-
Ce\, Cc2 = turbulence modeling pa- flux vector (Eq. (2)) netic energy
rameters Re = Reynolds number = boundary layer momentum
D, E = near-wall terms in k-e Re,,, Re, = turbulent Reynolds numbers thickness
equation ( R e v = (fky/v),Re,= (k2/ M/> M/ molecular, turbulent viscosity
fi,fi,fn,fw = near-wall damping func- vej) v = kinematic viscosity
tions Re» = momentum thickness Reyn- p = density
h = shape factor = 6*10 olds number ak, ac = Prandtl numbers in k and e equa-
H = heat transfer coefficient = S = coordinate along blade sur- tions
qJ{Tol - T„), W/m 2 K face (positive for SS and neg- Ty = Cartesian components of stress
k = turbulent kinetic energy ative for PS) tensor (Eq. (1))
= UiUj/2 SS = suction surface
K = acceleration factor = Subscripts
T = temperature
(vlUDdUJdS 7w„ = free-stream turbulence inten- e = boundary layer edge
sity is = isentropic
uT = friction velocity 1,2 = cascade inlet, exit condition
Uj = mean velocity component I, t = laminar, turbulent
U+ = U/uT o = stagnation condition
Ui = fluctuating velocity compo- w = wall
nent °° = inlet free stream

Journal of Turbomachinery OCTOBER 1997, Vol. 1 1 9 / 7 9 5

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


Table 2 Computed cases temperatures (T0] and Tw) are not representative of real engines,
but the ratio T„/T0] *» 0.7 is similar to those encountered in
Cases Pol( B «) T 0 l (k) R<V, (106) Mi,,, Tiu(%) T
w « modern jet engines.
MarklS 2.643 772 1.6 0.90 8.3 540 In the present computations, a 151 X 99 //-grid is employed.
Mark63 2.765 771 2.0 0.90 8.3 547 The predictions with a finer grid (171 X 121), were found to
3.907 719 2.5 0.91 8.3 597
be very close to those obtained from the 151 X 99 grid. About 25
Mark58
to 35 orthogonal grid points are distributed inside the boundary
C3X148 2.448 802 1.5 0.91 6.5 586
layers. The first grid point away from the wall is located at y''"
C3X158 2.435 808 1.5 0.90 8.3 590 as 1 in converged solutions. Both the original blade profiles and
C3X113 3.232 781 2.0 0.89 8.3 578 the blade profiles modified with cusped trailing edges have been
C3X156 4.007 781 2.5 0.89 8.3 656 used in the computations. It is found that the addition of a
cusped trailing edge greatly reduces the residue level in the
trailing edge region and required values of artificial dissipations,
without influencing the upstream boundary layer development.
to zero near the wall to avoid contamination of the solution by The modified blade profiles with cusped trailing edges for both
artificial dissipation. Mark II and C3X airfoils are used in the heat transfer computa-
An //-grid was used in the RK2D. The use of a combination tions presented below.
of algebraic and elliptic methods (Basson et al., 1993) makes Convergence criteria are taken as three decades drop of the
it possible to generate a smooth grid while maintaining strict residuals for rms density and turbulent kinetic energy. The dif-
control over grid spacing and orthogonality near the blade sur- ference between the mass flow rate at the inlet and at the exit
face. The grid orthogonality inside blade boundary layers is is typically less than 0.1 percent of the inlet mass flow rate.
highly desirable for implementation of low-Reynolds-number The computations of all these cases, by all three models, have
turbulence models. been carried out with identical grids, boundary conditions, arti-
At the inlet, the total pressure, total temperature, and flow ficial dissipation rates and convergence criteria for accurate
angle are specified. The left-running Riemann invariants are comparison and assessment. The converged solution from the
extrapolated from the interior of computation domain. The back CH model is used to initialize the computations with LB model
pressure is specified at the outlet for subsonic (axial) outlet and FLB model, respectively. It normally takes less than half
flow. an hour of CPU on the CRAY-YMP to obtain a convergent
For all three k- e models, k is zero at the wall. The e in the solution.
CH model is the isotropic component of the dissipation rate,
and thus should be zero at the wall. For the LB and FLB models, Blade Surface Pressure Distribution. It is known that the
the normal gradient of e is set to zero at the wall as suggested transition of turbine boundary layers is greatly affected by the
by Patel et al. (1985) and Fan et al. (1993). Constant values free-stream turbulence level and the surface pressure gradient.
of k and e are imposed at the inlet based on specified free- Increased free-stream turbulence promotes early transition,
stream turbulence intensity and length scale: while large flow acceleration tends to delay transition. The
Reynolds number and the Mach number (e.g., the shock wave
location) also influence the boundary layer transition.
Tu„ = ^ ^ (7) As shown in Fig. 1 (a), the predicted surface pressure distri-
butions are in excellent agreement with the data for both cas-
e= = k'J/U (8) cades. The pressure distributions are similar on the pressure
surfaces of these two airfoils. However, there is significant dif-
where the value of Tu* is available from experimental data. ference in the suction surface pressure distributions. A very
Following Hah (1989), the free-stream turbulence length scale strong adverse pressure gradient is clearly seen in the region
L» is assumed to be 1 percent of the pitch since the exact value from about 26 to 34 percent suction surface arc length (i.e.,
of Lc is not reported by Hylton et al. (1983). At the cascade ARC) on the Mark II airfoil, while the pressure distribution
exit, values for k, e and mean flow variables are extrapolated on the C3X airfoil suction surface exhibits only mild adverse
along j = constant grid lines. gradient.
The Navier-Stokes code with the three LRN k- e turbulence The predicted value of acceleration factor K (K =
models was validated by computing a flat-plate turbulent bound- (vlUl)dUJdS) is very helpful in an understanding of the
ary layer. The predictions of mean-velocity profile (w+ ~ y + ) boundary layer development, including transition and re-lami-
in the near-wall region with the three models were in good narization on the blade. Figure 1(b) shows the acceleration
agreement with the standard linear-log law. The predicted turbu- factors along both surfaces for Mark II and C3X airfoils at
lence quantities show good agreement with the experimental Re,.,,2 « 1.5 X 1 0 \ For values of K larger than 3.0 X 10~6
data (Patel et al., 1985). (relaminarization criterion), transition is suppressed for low
free-stream turbulence, and the turbulent boundary layer begins
Boundary Layer Code—TEXSTAN. The TEXSTAN code
to relaminarize. Over about half the length of the pressure sur-
(Crawford, 1985) is an extension of the well-known boundary
face, K is much larger than 3.0 X 10~6 for the Mark II airfoil.
layer code STAN5. A variety of LRN two-equation turbulence
In contrast, K is larger than the criterion only in a small region
models are available from this code.
close to the leading edge on the C3X pressure surface. On the
suction surface, Mark II airfoil exhibits a sharp spike of negative
Computation of Turbine Cascade Aerodynamics and acceleration at about 30 percent surface arc length. This spike
Heat Transfer corresponds to the large adverse pressure gradient observed in
that region (Fig. 1(a)). The C3X airfoil's acceleration factor
Test Cases and Numerical Details. The Mark II and C3X on the suction surface is seen to be close to zero except the
turbine guide vane cascades studied by Hylton et al. (1983) region near the leading edge (Fig. 1(b)).
and Nealy et al. (1984) were selected as the primary test cases
in the present work. The two cascades have close chord length Boundary Layer Development and Heat Transfer for
(around 14 cm) and turning angle (around 71 deg). The test Mark II and C3X. The predictions of heat transfer rate (nor-
conditions are given in Table 2 for these two cascades. malized by H,a, HKi = 1135 W/K/m 2 ) for MarklS (Table 2)
As shown in Table 2, the test conditions (Re;.yi2, M,Si2 and by both the Navier-Stokes and the boundary-layer methods are
Tua) are close to actual engine operations. The total and wall presented in Fig. 2(a). The LB k- e model used in the bound-

796 / Vol. 119, OCTOBER 1997 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


.^-icr.. -tr -trtJAI* confidence in the accuracy of the Cs prediction through the
\v Reynolds analogy. Indeed, the transition location indicated by
the variation in skin friction is consistent with that derived from
A/ '

7°' 1 1

d' a Data (Mark II)


I Computation (Mark II) V?
J 0 Data (C3x)
\\ooj * 0 ^>4-ni J*T'\
E — Computation (c3x)

Pressure Suction

1 1 l 5 0.6-
0.0
S/ARC

Fig. 1 ( a ) Surface static pressure distribution (for cases with M / s 2 = 0.90,


M a r k 1 5 and C3X158)

I —r
o.o 0.5 1.0
S/ARC
Relaminarization criterion
Mark II Fig. 2 ( a ) Heat transfer prediction for case M a r k 1 5 (M, s 2 = 0.90, Re, s2 :
C3x 1.6 x 1 0 6 , Tu„ = 8.3 percent)

20-

15-
1 —I 1
- 1 .0 -0.5 0.0
S/ARC

Fig. 1 (b) Acceleration factors for above cases ( M a r k 1 5 and C3X158)

ary-layer code is the same as that in the Navier-Stokes code.


The boundary layer prediction on the pressure side is in good -1
0.0 1.0
agreement with the measurement. The Navier-Stokes predic- S/ARC
tions also agree with the data well, the prediction by the CH
Fig. 2 ( b ) Skin friction ( C , ) prediction for case M a r k 1 5
k-e model being the best. The LB k-e model (in the Navier-
Stokes code) predicts a slightly early transition, which results
in an overprediction of the heat transfer to the rear of the surface.
The heat transfer over a large region near the midchord is under-
predicted by the FLB k-e model because this model predicts
a delayed transition. Some difference is observed between the
Navier-Stokes prediction and the boundary layer prediction
using the same LB k- e model. This difference can be attributed
to the difference in grid resolution, free-stream turbulence level
as well as the discretization scheme in these two codes.
On the suction side, the parabolic boundary layer computation
is terminated around S/ARC = 26 percent due to separation,
as shown in Fig. 2, while the elliptic Navier-Stokes computa-
tions with all the models have captured the sharp increase in
heat transfer due to separation-induced transition, with the pre-
diction by the FLB k-e model being the best. The transition Fig. 2 ( c ) Boundary layer shape factors [h) for case Mark15
onset location is observed to be coincident with the starting
location of strong adverse pressure gradient on the suction sur-
face (i.e., around 26 percent arc length, see Figs. 1(a) and
1(b)). The prediction by the CH k-e model exhibits an over-
shoot after the transition, which is due to the poor performance
of this model near the reattachment point. The major deficiency
of the CH k- e model lies in its use of y+ in the near wall
damping function /^. The wall shear stress is reduced to zero
at the separation point, which makes the inner variable y+ be-
come zero for any value of y. Thus the damping of eddy viscos-
ity is applied even to the outer region of the flow, which is
clearly incorrect.
The calculated skin friction coefficients (nondimensionalized
by local free-stream dynamic head) are plotted in Fig. 2(b).
Though experimental data are not available for the Cf, the agree-
ment between the heat transfer computation and data provides Fig. 2 ( d ) Boundary layer m o m e n t u m thicknesses for c a s e M a r k 1 5

Journal of Turbomachinery OCTOBER 1997, Vol. 1 1 9 / 7 9 7

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


the heat transfer distribution. From the Cf distribution on the
suction side, it is clear that the CH k-e model predicts the
largest separation bubble among the three models. On the pres-
sure side, the relative magnitudes of Cf computed by different
models are also consistent with those observed in the prediction
of heat transfer.
As can be seen from Fig. 2(c), the shape factor variation
computed by both the elliptic and parabolic methods follows
the heat transfer distribution. The shape factor predicted by the
FLB k-e model starts to decrease from 2.0 (corresponding to
laminar boundary layers) at about 75 percent pressure surface
arc length, which is exactly the location where the heat transfer
rate begins to increase (Fig. 2 ( a ) ) . Among all the models in — i — T
0.0 1.0
the Navier-Stokes code, the LB k- e model predicts the lowest
S/ARC
shape factor (about 1.4, corresponding to turbulent boundary
layers) and thus the highest heat transfer rate near the pressure Fig. 3(a) Heat transfer distribution along blade surfaces for case
surface trailing edge. On the suction side, the shape factors C3X158 (/Wfc2 = 0.90, Re„2 = 1.5 x 10 6 , T u . = 8.3 percent)
computed by the three models drop to around 1.35 near the
trailing edge.
In Fig. 2(d), the boundary layer momentum-thicknesses (9)
predicted by all the models at the rear of the pressure surface
are about 0.0004 pitch, which is much lower than the predicted
8 (about 0.002 pitch) at the rear of the suction surface. Also,
the pressure surface boundary layer thickness appears to grow
very slowly and even decreases in some region, primarily due
to the flow acceleration along the entire surface. Mee et al.
(1992) made similar observations from a detailed boundary
layer measurement of a transonic turbine cascade. The thin
boundary layer on the pressure surface made it necessary to
employ large number of grid points for proper near-wall flow
resolution.
The boundary layer properties and heat transfer for the case
C3X158 predicted by all three models are presented in Figs.
Fig. 3(b) Predicted boundary layer shape factors along blade surfaces
3 ( a ) , 3(b), 3(c), and 3(d). On the pressure surface, all three for case C3X158
models have captured transition with the prediction from the
LB k- e model being the closest to the data. Once again, the
FLB k- e model yields a delayed transition and thus underpre-
dicts the heat transfer over a large area on the surface. The C3X
airfoil exhibits a smooth transition on the suction surface since
the acceleration factor is close to zero except the leading edge
region (as seen in Fig. 1(b)). The transition onset predicted
by the CH k-e model is too close to the leading edge, and
the transition process appears to have been smeared out. The
tendency of the CH model in predicting a premature transition
has been observed by other researchers (Simoneau and Simon,
1993). The LB k- e model predicts a more distinct transition
process, but the transition onset is predicted to be earlier than
the data. The transition location predicted by the FLB k- e
model is in best agreement with the data.
As expected, the variation of boundary layer integral parame-
ters (see Figs. 3(b) and 3(c)) predicted by different models Fig. 3(c) Predicted momentum-thickness Reynolds number for case
are consistent with the heat transfer predictions from the same C3X158
model. As seen in Fig. 3(c), the computed Re9 near the trailing
edge reaches only around 750, which is far below a value of
2000 usually considered as the lower limit for fully turbulent 25-
boundary layers. Thus, from the predicted heat transfer, shape - r^*""" "

factor, and Re 9 , it may be concluded that the boundary layer 20-


— linear & log-law

remains transitional along the entire pressure surface. On the S/ARC=14%


S/ARC=22%
y " .-•' * "y'J^^"
•'i^^*^
suction surface, the predicted Re e is around 2200 near the trail-
15-
- - • S/ARC=36%
S/ARC=51%
/.---
,/..-j&
ing edge. The streamwise evolution of suction surface boundary
layer velocity profile (u + ~ y + ) predicted by the LB k- e model
is shown in Fig. 3(d). The prediction provides a clear simula- 10- y"^
tion of the transition process in which the velocity profile in
outer layer changes from the type of a laminar boundary layer 5 -
to that of a fully turbulent boundary layer (log-law). Thus, the
computed heat transfer, shape factor, Refl and u+ ~ y+ (Figs. I
I I I ) 1 I 1 1 I 1 . .

3(a),3(b),3(c), and3(<i)) all indicate that the suction surface


boundary layer has become fully turbulent near the trailing edge.
From Figs. 2 and 3, it maybe concluded that the k-e model Fig. 3(d) Predicted velocity profile (by Lam-Bremhorst model) along
of Lam-Bremhorst (1981) provides the best prediction of tran- suction surface (case C3X158)

798 / Vol. 119, OCTOBER 1997 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


1.2- o Data (Re=1.6x10)
A Data (Re=2.0x10*)
>
o
A
Data (Mark15)
Data (03X158)
1.0-
D Data (Re=2.5x10°) fa-< a
;;.., LB comp. (Re=1.6x10 )
LB (Mark15)
LB (C3X158) " \ \ LB comp. (Re=2.0x1 o")
^ta*g A
0.8-
N ^ LB comp. (Ra=2.5x10 )
o ° o * 35
OiSnfT
5" 0 . 6 - XD ef.\ • '|R ° V&5~
X $ O <B
O
0.4-

0.2-
*
0.0-
i i i 1 1
1 - I —
1
-1 .0 0.0 0.5 -r 0.0
1.0 S/ARC
S/ARO

Fig. 4(a) Effects of pressure gradient on heat transfer distribution Fig. 5 Effects of Reynolds number on Mark II heat transfer distribution
(computation with Lam-Bremhorst k-e model)
(Mark15 and C3X158)

however, the Reynolds number effect on the heat transfer distri-


bution is reflected in both the transition onset location and the
general level of heat transfer (see Fig. 6). The higher the Reyn-
olds number, the earlier is the transition and thus the higher the
heat transfer rate. The LB k- e model captures this trend quite
well, especially on the pressure surface.
Free-Stream Turbulence Intensity Effect. In the experi-
ments carried out by Hylton et al. (1983), the free-stream turbu-
lence intensity 7w„ was varied from 6.5 to 8.3 percent for the
C3X airfoil. Figure 7 shows the effect of free-stream turbulence
intensity on the heat transfer level. The computed results (by
- 1 —
-0.5 ~r- - 1 —
0.6
0.0

Fig. 4(b) Effect of pressure gradient on momentum thickness variation


(Mark15 and C3X158)

sitions on these two airfoils among the three k- e models evalu-


ated. Sieger et al. (1993) reached a similar conclusion after a
detailed evaluation of various LRN k- e models for predicting
bypass transition, using a boundary layer code.
Pressure Gradient Effect. Since the major difference be-
tween the Mark II and the C3X airfoils lies in the suction surface
pressure distribution, a comparison of heat transfer rates on this
two airfoils is made in Fig. 4 ( a ) . The test conditions (Re, Mach — i —
0.0
number, Tu„) are the same for the two cases. On the pressure
surface, the C3X airfoil exhibits an earlier transition than the
Mark II airfoil, which is due to the fact that the acceleration Fig. 6 Effect of Reynolds number on C3X heat transfer distribution
(computation with Lam-Bremhorst k-e model)
factor of the C3X airfoil is lower than that of the Mark II airfoil
along the entire pressure surface (Fig. 1(b)). On the suction
surface, the increase in local heat transfer caused by the separa-
tion-induced transition (on Mark II) is much sharper than that
caused by the nominal bypass transition (on C3X). Similar
behavior in the variation of the momentum thickness can be
seen in Fig. 4(b).

Reynolds Number Effect. The effect of Reynolds number


on the airfoil heat transfer distribution is shown in Figs. 5 and
6 for the Mark II and C3X airfoils, respectively. The predictions
by the LB k-e model generally capture the increase in heat
transfer rate with increasing Reynolds numbers for both airfoils.
As discussed in Hylton et al. (1983) and Nealy et al. (1984),
in the case of Mark II airfoil, the Reynolds number effect (at
a fixed M(.,2) appears to be largely reflected in a shift in the I
0.5
o.o
heat transfer level rather than in heat transfer distribution, since S/ARO
the transition apparently occurs at a fixed location (i.e., around
26 percent arc length, Fig. 5). The transition predicted by the Fig. 7 Free-stream turbulence intensity effect on heat transfer distribu-
tion along blade surfaces for case C3X148 and C3X158 (/W,s2 = 0.90, Refa2
LB k- e model is independent of Reynolds number. This is in = 1.5*10 6 , 7u„ = 6.5 percent for C3X148 and Tu„ = 8.3 percent for
good agreement with the measurement. For the C3X airfoil, C3X158)

Journal of Turbomachinery OCTOBER 1997, Vol. 1 1 9 / 7 9 9

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


stagnation region. However, Eq. (9) will inevitably provide a
very high level of turbulence production due to the existance
O Data (Mark15)
FLB(L=1% pitch) of the square of large strains multiplied by a large eddy viscos-
FLB(L=3% pitch) ity. The inaccurately amplified turbulent kinetic energy in the
-— FLB(L=5% pitch)
leading edge region is convected downstream, along both sur-
faces. The velocity gradient (and Pk in Eq. (9)) downstream
; o.6' of the leading edge on the suction side is much higher than that
on the pressure side, thus the free-stream turbulence level near
the suction surface as computed by the k- e models is signifi-
cantly higher than that near the pressure surface, resulting in
significant overpredictions of heat transfer on the suction sur-
face near the leading edge (Figs. 8 and 9). For an accurate
—I 1 - T prediction of free-stream turbulence in the stagnation region, it
-0.5 0.0 1.0
is thus necessary to employ anisotropic models, such as the
S/ARO
low-Reynolds-number Reynolds stress models.
Fig. 8 Length scale effect on heat transfer distribution for case Mark15
(computation with Fan-Lakshminarayana-Barnett k-e model)
Concluding Remarks
The numerical simulations discussed in this paper indicate
the CH k-e model) are almost identical for the two cases that the turbine blade heat transfer, under conditions close to
(C3X148 and C3X158). The predictions from the other two real engine operating conditions (high free-stream turbulence,
models (not shown) indicate similar features. The experimental high Re, and favorable-to-adverse pressure gradient), can be
data indicates a slightly higher level of heat transfer on both predicted with engineering accuracy by the three low-Reynolds-
surfaces under higher free-stream turbulence intensity (i.e., case number k-e models developed by Chien (1982), Lam-Brem-
C3X158). horst (1981), and Fan et al. (1993), respectively. The Navier-
Stokes predictions with various low-Reynolds-number k- e
models capture the location of separation-induced transition
Free-Stream Turbulence Length Scale Effect. As stated very well. Both the data and computations indicate that the
earlier, the free-stream turbulence length scale L„ was specified separation-induced transition results in much sharper increase
as 1 percent of the pitch at the computational inlet (about 0.8 in the local heat transfer rate than the nominal bypass transition.
axial chord upstream of the leading edge), since no data are The effects of Reynolds number have been captured well by all
available for this parameter. The effect of this length scale on the models. It has also been shown that the Navier-Stokes
the airfoil heat transfer distribution is examined in Figs. 8 and procedure provides accurate predictions of both heat transfer
9 for the Mark II and C3X airfoils, respectively. Increasing the and boundary layer aerodynamic parameters.
free-stream length scale is equivalent to increasing the free-
stream turbulence intensity, since larger length scale means
lower decay rate of free-stream turbulence (Eq. (8)) and hence
higher turbulence level at the blade leading edge. Indeed, the
predicted heat transfer rate increases with the specified value
of Leo, as seen in Fig. 8, Figs. 9(a) and 9(b). Predictions from O Data (03X158)
different models exhibit quite similar trends. CH ( L = 1 % pitch)
CH (L=3% pitch)
However, it should be noted that the predicted heat transfer CH (L=5% pitch)

on the suction surface is much more sensitive to the variation


in L„ than that on the pressure surface. This is essentially due
to the isotropic character of the present k-e models, which
provide a poor representation of normal stress components even
in a flat plate boundary layer. It is known that isotropic models
are not accurate for stagnation flow, where the production of
turbulence is dominated by normal stress components. As
shown by Taulbee et al. (1989), the turbulence level along the 1 - 1 — - T
0.0 0.5 1.0
stagnation streamline is overpredicted considerably by the k- e S/ARC
model, but is predicted accurately with the Reynolds stress
model. This can be understood by a comparison of the produc- Fig. 9(a) Length scale effect on heat transfer distribution for case
tion terms Pk. For two-dimensional incompressible flows, Pk C3X158 (computation with Chien k-e model)
arising from an isotropic eddy-viscosity model and a Reynolds
stress model are given by
o Data (C3X158)
LB ( L = 1 % pitch)
LB (L=3% pitch)

2,/^y + 2,/^y +v m + ^ (9)


— - LB (L=5% pitch)

\dx) \dy) \dy 8x

^.dU _ dV 3U dV\
Pk = —uu — w —— uv I 1 (10)
ox ay
dy dx I

The large velocity gradient (i.e., dUldx) during the decelera-


tion of the flow toward the stagnation point makes the first two -~1 —I— I —r
terms (dV/dy = -dU/dx due to continuity equation) to be -0.5 0.0 0.5 1.0
dominant in Pk. Since uu does not differ greatly from vv, the
first two terms in Eq. (10) (i.e., (-Tm + vv)dU/dx) can still Fig. 9(b) Length scale effect on heat transfer distribution for case
yield a nominal level of turbulent energy production in the C3X158 (computation with Lam-Bremhorst k-e model)

800 / Vol. 119, OCTOBER 1997 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo


The Lam-Bremhorst k-e model provides relatively better Fan, S., Lakshminarayana, B., and Barnett, M., 1993, "Low-Reynolds-Number
k- e Model for Unsteady Turbulent Boundary-Layer Flows," A1AA Journal, Vol.
predictions of the transition and the heat transfer than the other 32, No. 10, pp. 1777-1784.
two models for the cases studied. While performing well in Fan, S., and Lakshminarayana, B., 1996, "Computation and Simulation of
fully turbulent boundary layers, Chien's k-e model tends to Wake-Generated Unsteady Pressure and Boundary-Layers in Cascades: Parts 1 &
predict a premature transition and smear out the transition pro- 2," ASME JOURNAL OF TURBOMACHINERY, Vol. 118, pp. 96-122.
cess. This model is also not accurate for separated flows. The Hah, C , 1989, "Numerical Study of Three-Dimensional Flow and Heat Trans-
fer Near the Endwall of a Turbine Blade Row," AIAA Paper No. 89-1689.
Fan-Lakshminarayana-Barnett k-e model provides the best Hylton, L. D„ Mihelc, M. S., Turner, E. R., Nealy, D. A., and York, R. E„
prediction of separation-induced transition, but tends to predict 1983, "Analytical and Experimental Evaluation of the Heat Transfer Distribution
delayed transitions for accelerating boundary layers on the pres- Over the Surface of Turbine Vanes," NASA CR-168015, May.
sure surface. Kunz, R., and Lakshminarayana, B., 1992, "Explicit Navier-Stokes Computa-
tion of Cascade Flows Using the k-e Model," AIAA Journal, Vol. 30, No. 1.
Detailed experiments should be performed to examine the Lam, C , and Bremhorst, K., 1981, "Modified Form of the k-e Model for
variation of free-stream turbulence through the passage, particu- Predicting Wall Turbulence," ASME Journal of Fluids Engineering, Vol. 103,
larly near the stagnation point and accelerating region on the p. 456.
suction surface. To improve the predictions of convective heat Luo, J., and Lakshminarayana, B., 1995, "Navier-Stokes Analysis of Turbine
Flowfield and Heat Transfer," J. of Propulsion and Power, Vol. 11, pp. 2 2 1 -
transfer near the stagnation point, low-Reynolds-number aniso- 229.
tropic turbulence models such as a low-Reynolds-number Reyn- Mayle, R. E., 1991, "The Role of Laminar-Turbulent Transition in Gas-Turbine
olds stress model should be used. Engines," ASME JOURNAL OF TURBOMACHINERY, Vol. 113, pp. 509-537.
Mee, D. J., Baines, N. C , and Oldfield, M. L. G., 1992, "Detailed Boundary
Layer Measurements on a Transonic Turbine Cascade," ASME JOURNAL OF
Acknowledgments TURBOMACHINERY, Vol. 114, pp. 163-172.
Nealy, D. A., Mihelc, M. S., Hylton, L. D„ and Gladden, H. J., 1984, "Measure-
This work was supported by the U.S. National Aeronautics ments of Heat Transfer Distribution Over the Surfaces of Highly Loaded Turbine
and Space Administration through contract NAS 8-38867 moni- Nozzle Guide Vanes," ASME Journal of Engineering for Gas Turbines and
tored by Lisa Griffin of Marshall Space Flight Center (MSFC). Power, Vol. 106, pp. 149-158.
Patel, V. C , Rodi, W., and Scheuerer, G„ 1985, "Turbulence Models for Near-
The authors wish to acknowledge NASA for providing super- Wall and Low-Reynolds-Number Flows," AIAA Journal, Vol. 23, No. 9, pp.
computer resources at the NASA Ames research center. The 1308-1319.
National Science Foundation grant (NSF Grant INT 90-15984) Savill, A. M., 1993, "Some Recent Progress in the Turbulence Modeling of
enabled interaction with the turbulence group at Ecole centrale By-Pass Transition," Near-Wall Turbulent Flows, R. M. C. Soetal., eds., Elsevier
de Lyon (France). Science Publishers.
Schmidt, R., and Patankar, S., 1992, "Simulating Boundary Layer Transition
With Low-Reynolds-Number k- c Turbulence Models," ASME JOURNAL OF TUR-
BOMACHINERY, Vol. 114, pp. 10-17.
References Sieger, K„ Schulz, A., Crawford, M. E., and Wittig, S„ 1993, "An Evaluation
Amei'i, A. A., and Arnone, A., 1994, "Prediction of Turbine Blade Passage of Low-Reynolds-Number k- c Turbulence Models for Predicting Transition Un-
Heat Transfer Using a Zonal and a Two-Equation Turbulence Model," ASME der the Influence of Free-Stream Turbulence and Pressure Gradient," Engr. Turbu-
Paper No. 94-GT-122. lence Modelling and Experiments, W. Rodi and F. Martelli, eds., Elsevier Science
Basson, A. H., Kunz, R. F., and Lakshminarayana, B., 1993, "Grid Generation Publishers.
for Three-Dimensional Turbomachinery Geometries Including Tip Clearance," Simoneau, R. J., and Simon, F. F., 1993, "Progress Towards Understanding
A1AA Journal of Propulsion and Power, Vol. 9, p. 59. and Predicting Heat Transfer in the Turbine Gas Path," Inter. J. of Heat & Fluid
Chien, K. Y., 1982, "Prediction of Channel and Boundary-Layer Flows With Flow, Vol. 14, pp. 106-128.
a Low-Reynolds-Number Turbulence Model," A1AA Journal, Vol. 20, No. 1. Taulbee, D. B„ Tran, L. T., and Dunn, M. G., 1989, "Stagnation Point and
Crawford, M., 1985, TEXSTAN program, Dept. of Mech. Engr., Univ. of Texas Surface Heat Transfer for a Turbine Stage: Prediction and Comparison With
at Austin. Data," ASME JOURNAL OF TURBOMACHINERY, Vol. I l l , pp. 28-35.

Journal of Turbomachinery OCTOBER 1997, Vol. 1 1 9 / 8 0 1

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/28663/ on 06/10/2017 Terms of Use: http://www.asme.org/abo

You might also like