three-dimensional boundary layer in a turbine blade passage

10
Three-Dimensional Boundary Layer in a Turbine Blade Passage Stephen Lynch Pennsylvania State University, University Park, Pennsylvania 16802 DOI: 10.2514/1.B36232 The boundary layer on the end wall of a turbine blade cascade is subject to cross-stream pressure gradients in the blade passage, which generate a cross-stream velocity component to make it three dimensional. This distorts the turbulence relative to a two-dimensional boundary layer and impacts the end wall heat transfer. This study presents measurements of the three-dimensional boundary layer in a turbine cascade obtained with a laser Doppler velocimeter. In addition, two types of Reynolds-averaged NavierStokes models are compared to the measurements: the Shear Stress Transport k ω model using the isotropic eddy viscosity assumption and a Reynolds stress model that allows for anisotropy of the Reynolds stress. Neither model fully captures the complexity of the three-dimensional boundary layer in a turbine blade passage, particularly for turbulence associated with the cross-stream flow and for the highly accelerated three-dimensional boundary layer at the passage exit. Measurements at the passage exit indicate a very thin boundary layer with laminarlike qualities. Because of the boundary-layer complexity, end wall heat transfer is not well predicted toward the pressure side and the exit of the blade passage. Nomenclature C ax = axial chord of blade, m C f = friction coefficient, τ w 12U 2 edge D = diameter of laser Doppler velocimeter measure- ment volume, m F = SST k ω turbulence model blending function H = boundary-layer shape factor, δ θ h = heat transfer coefficient, q 00 T wall T in , Wm 2k k = turbulent kinetic energy, 12u 0 u 0 v 0 v 0 w 0 w 0 , m 2 s 2 L = length of laser Doppler velocimeter measurement volume, m P = blade pitch, m, or static pressure, Pa Pr t = turbulent Prandtl number q 00 = wall heat flux, Wm 2 Re θ = momentum thickness Reynolds number, θU edge v S = blade span, m or mean strain rate tensor, 1s St = Stanton number, hρC p U in or hρC p U edge SS = suction side s = limiting streamline direction, m T = temperature, K Tu = freestream turbulence level, u 0 U edge U, V, W = mean velocity components in cascade coordinate system, m/s U edge = magnitude of local freestream velocity at boundary-layer edge, m/s U exit = magnitude of exit streamwise velocity, m/s U in = magnitude of inlet streamwise velocity, m/s u τ = friction velocity, τ w ρ p , m/s u 0 , v 0 , w 0 = fluctuating components in cascade coordinate system, m/s X, Y, Z = cascade coordinates, where X is blade axial direction, m/s α = thermal diffusivity, m 2 s or turbulence model constant β = mean flow angle, deg, or turbulence model constant, deg δ ij = Kronecker delta δ 99 = boundary-layer thickness (99%), m δ = displacement thickness, m θ = momentum thickness, m μ t = turbulent viscosity, N s m 2 ν = kinematic viscosity, m 2 s ρ = density, kgm 3 σ = turbulence model constant τ w = wall shear stress, Pa ω = specific dissipation, 1s Subscript FS = based on local freestream direction (determined at midspan) Superscripts = quantity based on inner coordinates (e.g., u is equal to uu τ ) = time-averaged quantity (for correlations of fluctuating velocities). I. Introduction T HREE-DIMENSIONAL boundary layers are found in many practical situations in which a cross-stream pressure gradient exists to turn the flow, such as in curved channels, around the base of bridge piers, and within axial turbomachine passages. For a given cross-stream pressure gradient, the radius of curvature of the low- momentum streamlines very near the end wall is smaller than for the higher-momentum fluid outside the boundary layer, and the end wall boundary layer develops a component of velocity in the cross-stream direction. This is Prandtls secondary flow of the first kind and results in a skewed boundary layer that impacts the development of turbulence and associated convective heat transfer relative to a two- dimensional flow. For an axial gas turbine, high levels of flow turning are desirable to extract maximum work from the high-enthalpy gases. However, this results in the development of significant crossflow in the end wall boundary layer. This crossflow implies mean streamwise vorticity in the near-wall flow, which becomes significant as it progresses through the axial turbine passage. Additional streamwise vorticity, generated at the leading edge of the turbine airfoil due to the rollup of the inlet boundary layer upstream of the airfoil, is thought to merge with the secondary flow to create a large feature known as the passage vortex. This vortex is responsible for high heat transfer coefficients on the turbine end wall and airfoil surfaces, as well as increased Presented as Paper 2016-1891 at the 54th AIAA Aerospace Sciences Meeting, San Diego, CA, 48 January 2016; received 3 March 2016; revision received 27 September 2016; accepted for publication 11 October 2016; published online 19 January 2017. Copyright © 2016 by Stephen Lynch. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. All requests for copying and permission to reprint should be submitted to CCC at www.copyright.com; employ the ISSN 0748-4658 (print) or 1533-3876 (online) to initiate your request. See also AIAA Rights and Permissions www.aiaa.org/randp. *Assistant Professor, Department of Mechanical and Nuclear Engineering. Senior Member AIAA. 954 JOURNAL OF PROPULSION AND POWER Vol. 33, No. 4, JulyAugust 2017 Downloaded by PENNSYLVANIA STATE UNIVERSITY on April 13, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.B36232

Upload: others

Post on 27-Feb-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Three-Dimensional Boundary Layer in a Turbine Blade Passage

Three-Dimensional Boundary Layer in a Turbine Blade Passage

Stephen Lynch∗

Pennsylvania State University, University Park, Pennsylvania 16802

DOI: 10.2514/1.B36232

The boundary layer on the end wall of a turbine blade cascade is subject to cross-stream pressure gradients in the

blade passage, which generate a cross-stream velocity component to make it three dimensional. This distorts the

turbulence relative to a two-dimensional boundary layer and impacts the end wall heat transfer. This study presents

measurements of the three-dimensional boundary layer in a turbine cascade obtained with a laser Doppler

velocimeter. In addition, two types of Reynolds-averaged Navier–Stokes models are compared to the measurements:

the Shear Stress Transport k − ω model using the isotropic eddy viscosity assumption and a Reynolds stress model

that allows for anisotropyof theReynolds stress. Neithermodel fully captures the complexity of the three-dimensional

boundary layer in a turbine blade passage, particularly for turbulence associated with the cross-stream flow and for

the highly accelerated three-dimensional boundary layer at the passage exit. Measurements at the passage exit

indicate a very thin boundary layer with laminarlike qualities. Because of the boundary-layer complexity, end wall

heat transfer is not well predicted toward the pressure side and the exit of the blade passage.

Nomenclature

Cax = axial chord of blade, mCf = friction coefficient, τw∕�1∕2�U2

edge

D = diameter of laser Doppler velocimeter measure-ment volume, m

F = SST k − ω turbulence model blending functionH = boundary-layer shape factor, δ�∕θh = heat transfer coefficient,q 0 0∕�Twall − Tin�,W∕m2−k

k = turbulent kinetic energy, �1∕2���u 0u 0� � �v 0v 0���w 0w 0��, m2∕s2

L = length of laser Doppler velocimeter measurementvolume, m

P = blade pitch, m, or static pressure, PaPrt = turbulent Prandtl numberq 0 0 = wall heat flux, W∕m2

Reθ = momentum thickness Reynolds number, θUedge∕vS = blade span, m or mean strain rate tensor, 1∕sSt = Stanton number, �h∕ρCpUin� or �h∕ρCpUedge�SS = suction sides = limiting streamline direction, mT = temperature, KTu = freestream turbulence level, u 0∕Uedge

U, V,W = mean velocity components in cascade coordinatesystem, m/s

Uedge = magnitude of local freestream velocity atboundary-layer edge, m/s

Uexit = magnitude of exit streamwise velocity, m/sUin = magnitude of inlet streamwise velocity, m/suτ = friction velocity,

�����������τw∕ρ

p, m/s

u 0, v 0, w 0 = fluctuating components in cascade coordinatesystem, m/s

X, Y, Z = cascade coordinates, where X is blade axialdirection, m/s

α = thermal diffusivity, m2∕s or turbulence modelconstant

β = mean flow angle, deg, or turbulence modelconstant, deg

δij = Kronecker deltaδ99 = boundary-layer thickness (99%), mδ� = displacement thickness, mθ = momentum thickness, mμt = turbulent viscosity, N−s∕m2

ν = kinematic viscosity, m2∕sρ = density, kg∕m3

σ = turbulence model constantτw = wall shear stress, Paω = specific dissipation, 1∕s

Subscript

FS = based on local freestream direction (determined atmidspan)

Superscripts

� = quantity based on inner coordinates (e.g., u� isequal to u∕uτ)

�� = time-averaged quantity (for correlations offluctuating velocities).

I. Introduction

T HREE-DIMENSIONAL boundary layers are found in manypractical situations in which a cross-stream pressure gradient

exists to turn the flow, such as in curved channels, around the base ofbridge piers, and within axial turbomachine passages. For a givencross-stream pressure gradient, the radius of curvature of the low-momentum streamlines very near the end wall is smaller than for thehigher-momentum fluid outside the boundary layer, and the end wallboundary layer develops a component of velocity in the cross-streamdirection. This is Prandtl’s secondary flow of the first kind and resultsin a skewed boundary layer that impacts the development ofturbulence and associated convective heat transfer relative to a two-dimensional flow.For an axial gas turbine, high levels of flow turning are desirable to

extract maximum work from the high-enthalpy gases. However, thisresults in the development of significant crossflow in the end wallboundary layer. This crossflow implies mean streamwise vorticity inthe near-wall flow, which becomes significant as it progressesthrough the axial turbine passage. Additional streamwise vorticity,generated at the leading edge of the turbine airfoil due to the rollup ofthe inlet boundary layer upstream of the airfoil, is thought to mergewith the secondary flow to create a large feature known as the passagevortex. This vortex is responsible for high heat transfer coefficientson the turbine end wall and airfoil surfaces, as well as increased

Presented as Paper 2016-1891 at the 54th AIAA Aerospace SciencesMeeting, San Diego, CA, 4–8 January 2016; received 3March 2016; revisionreceived 27 September 2016; accepted for publication 11 October 2016;published online 19 January 2017. Copyright © 2016 by Stephen Lynch.Published by the American Institute of Aeronautics and Astronautics, Inc.,with permission. All requests for copying and permission to reprint should besubmitted to CCC at www.copyright.com; employ the ISSN 0748-4658(print) or 1533-3876 (online) to initiate your request. See also AIAA Rightsand Permissions www.aiaa.org/randp.

*Assistant Professor, Department ofMechanical andNuclear Engineering.Senior Member AIAA.

954

JOURNAL OF PROPULSION AND POWER

Vol. 33, No. 4, July–August 2017

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 2: Three-Dimensional Boundary Layer in a Turbine Blade Passage

aerodynamic loss due to increased mixing of the surface boundary-layer fluid with higher-velocity mainstream fluid.The development of the turbine passage secondary flow away from

end wall and airfoil surfaces has been fairly well captured by design-level predictive codes. However, the accurate prediction of surfacequantities (wall shear stress and heat transfer coefficients) is stilllacking, due to an incomplete understanding of the influence ofthree dimensionality in the end wall boundary layer. Very fewmeasurements of this region exist to date, due to the difficulty ofacquiring measurements within a turbine blade passage. This paperpresents measurements and predictions of the turbine end wallboundary layer in the forward portion of the blade passage tohighlight regions in which the effects of boundary-layer threedimensionality are important to capture.

II. Previous Studies

The end wall boundary layer in an axial turbine is not only subjectto large cross-stream pressure gradients but also to streamwisepressure gradients due to flow acceleration through the passage.Also,as mentioned in the Introduction, a passage vortex develops whenthe passage secondary flow merges with the streamwise vorticitygenerated by the inlet boundary-layer separation at the airfoil leadingedge. Various studies have examined individual contributions ofthese effects, as well as their combined influence, on the nature of thethree-dimensional turbulent boundary layer (3DTBL).A 3DTBL can be generated by transverse pressure gradients

(pressure-driven) or by shear (shear-driven), such as on a rotatingdisk. The resulting cross-stream flow manifests as mean streamwisevorticity with a scale no larger than the size of the boundary layer.Several researchers have found significant structural changes to theturbulence in a 3DTBL that distinguish it from a two-dimensionalboundary layer. Anderson and Eaton [1], Coleman et al. [2], Moinet al. [3], Schwarz andBradshaw [4], and Flack and Johnston [5] haveall reported that the vector of Reynolds shear stress (with componentsformed by thewall normal-to-streamwise correlated fluctuations andthe wall normal-to-cross-stream correlated fluctuations) lags themean strain rate vector for a pressure-driven 3DTBL, implyinganisotropic eddy viscosity. A review of several three-dimensionalboundary-layer data sets by Johnston and Flack [6] showed that theratio of the Reynolds shear stress magnitude to the turbulent kineticenergy tended to decrease as the total skew angle between thefreestream and the wall shear increased, implying reduced ability ofthe boundary layer at generating Reynolds shear stresses. Hydrogenbubble visualization by Flack [7] suggested that this effect may bedue to mean three dimensionality acting to stabilize the turbulentstructures generated in the boundary layer.In a 3DTBL subject to alternating cross-stream pressure gradients,

the crossflow can reverse direction. Bruns et al. [8] studied thissituation in an S-shaped bend, and Olcmen and Simpson [9]examined this for the flow around awing/body junction. Olcmen andSimpson [9] determined that the ratio of cross-stream to streamwiseeddy viscosity in the three-dimensional boundary layer scaled best ina coordinate system aligned with the wall shear direction, but laterwork (Olcmen and Simpson [10]) found these coordinates did notuniversally collapse the data of others for several three-dimensionalturbulent boundary-layer data sets.Onlya fewstudies have considered convective heat transfer effects in

three-dimensional turbulent boundary layers. Time-mean heat transfermeasurements byAbrahamson and Eaton [11] and Lewis and Simpson[12] agreed with two-dimensional correlations of Stanton number andenthalpy thickness Reynolds number, if the Reynolds number wasbased on the magnitude of a vector formulation for enthalpy thickness.Increased cross-stream pressure gradients reduced the measured time-mean heat transfer coefficients relative to a two-dimensional boundarylayer, but Lewis and Simpson [12] found that profiles of temperaturefluctuations were not affected by three dimensionality.Fundamental studies of a streamwise vortex embedded in a

boundary layer have indicated that the vortex can have a significantimpact on end wall heat transfer; although it should be noted that thestreamwise vortex along the endwall in a turbine passage is generally

larger than the boundary layer. Eibeck and Eaton [13] found thatincreasing levels of circulation produced by a delta wing in a two-dimensional boundary layer caused local Stanton numbers toasymptote to approximately 24% augmentation relative to a flat plateStanton number in the downwash region of the vortex. Heat transfereffects of the streamwise vortexwere attributed solely to its distortionof the mean flow and not to turbulence enhancement, although noturbulence measurements were provided. However, fluctuatingvelocity and temperature measurements by Wroblewski and Eibeck[14] showed that turbulent transport of heat was larger than turbulenttransport of momentum in the downwash region of a streamwisevortex, suggesting that heat transfer enhancement mechanisms aremore complex than only the distortion of the mean flow.The combined effects of a streamwise vortex and a pressure-driven

three-dimensional boundary layer were considered by Shizawa andEaton [15]. A streamwise vortex with near-wall induced velocity inthe same direction as the crossflow reduced the strength of ejectionevents. They also found that the three dimensionality of the flowdissipated the streamwise vortexmore quickly than a similar vortex ina two-dimensional flow.The large-scale features of the endwall flowwithin a turbine airfoil

passage have been investigated in detail by several researchers.Langston [16], Sharma andButler [17], andWang et al. [18] proposedmodels of the flow in a turbine airfoil passage. Common elements ofall models include a leading-edge horseshoe vortex with pressure-and suction-side legs as well as a large passage vortex and severalsmaller corner vortices. The horseshoe vortex at the leading edgedevelops due to the separation of the incoming end wall boundarylayer, and its legs progress around the airfoil. Because of theasymmetry of the turning flow, the pressure-side leg is generallylarger and eventuallymergeswith the secondary flow that develops inthe passage due to flow turning. Langston et al. [19], Kang and Thole[20], Harrison [21], Gregory-Smith et al. [22], and Knezevici et al.[23] all report measurements of these features and their contributionto high total pressure loss and high associated end wall heat transfer.Turbulence measurements by Kang and Thole [20] showed the wall-normal component was larger than either the streamwise or cross-stream component and that the turbulence intensity in the passagevortex core was around 26%, based on the inlet velocity.Characterization of the end wall boundary layer in an airfoil

cascade has only been attempted in a few studies. Measurements inthe passage by Langston et al. [19] and Harrison [21] suggest a thin,possibly laminar boundary layer on the end wall downstream of theinlet boundary-layer separation. Harrison [21] found that hot-filmsignals in that region also exhibited much lower rms values ascompared to the upstream turbulent boundary layer. More recently,hot film signals presented byVera et al. [24] indicated low amplitudesnear the pressure side, suggesting laminarlike flow. Signals increasedin amplitude and unsteadiness from the pressure side to the suctionside of the adjacent airfoil.Recent work by the author of this paper (Lynch andThole [25]) has

investigated the nature of the 3DTBL in the forward part of a turbineblade passage. Measurements of the mean flow indicated significantlevels of skew (defined as the angle between the endwall shear vectorand the freestream mean flow vector) of greater than 70 deg, whichare higher than any reported in the literature on 3DTBLs. This paperreports the companion turbulent quantities for the 3DTBL in a turbineblade passage as well as comparisons to an isotropic eddy viscositymodel (SST k − ω by Menter [26]) and a Reynolds stress model(RSM) (specifically the low-Reynolds Stress-Omega model byWilcox [27]), which allows for Reynolds stress anisotropy.

III. Description of Experiment

Details of the experimental setup are given by Lynch et al.[25,28,29], and, so only a brief treatment will be presented here.Measurements of the end wall boundary layer in a turbine bladepassage were obtained in a large-scale cascade comprised of low-pressure turbine (LPT) airfoils. The cascade was connected to aclosed-loop recirculating wind tunnel as shown in Fig. 1. A bi-planebar grid was located upstream of the cascade to provide a nominal

LYNCH 955

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 3: Three-Dimensional Boundary Layer in a Turbine Blade Passage

freestream turbulence level of 4% at the cascade inlet plane, which isrepresentative of the operating environment of a LPT blade. The inletvelocity boundary layer was measured 2.85 axial chords upstream ofthe cascade, and its various descriptive quantities are shown inTable 1. The incoming temperature profilewas uniform until the startof the endwall heating section. An endwall heater supplied a uniformheat flux of 1000 W∕m2 to the central four blade passages, startingfrom 3.4Cax upstream of the airfoils to 1.0Cax downstream.The turbine blade cascade contained seven LPT airfoils based on

the Pack-B design, which has been used in several studies of flowcontrol (Bons et al. [30] andMcAuliffe and Sjolander [31]), end wallsecondary flow (Knezevici et al. [23] and Praisner et al. [32]), and endwall heat transfer (Lynch et al. [29] andMensch and Thole [33]). Thescale of the cascade was 8.6 times the engine scale to allow for highmeasurement resolution. The cascade was operated at an engine-relevant exit Reynolds number of 200,000. Because of the large scaleand the capability of the low-speed wind tunnel, the Mach numberwas not matched to engine conditions. Details of the airfoil geometryand operating conditions are given in Table 2, and themeasured bladesurface static pressure at midspan (adapted from Lynch et al. [29]) isgiven in Fig. 2. The predicted blade static pressure distribution inFig. 2 from the SST k − ω model agreed very well with themeasurements. Although not shown here, the other prediction modelsinvestigated also agreed well with the midspan static pressure.Measurements of all three velocity components and turbulence

quantities were obtained with a three-component laser Dopplervelocimeter (LDV), mounted above the cascade as shown in Fig. 3.The top end wall of the cascade was constructed of glass.Measurements were obtained on the bottom end wall of the cascade,which was painted flat black to minimize stray reflections. An argon-ion laser supplied the green, blue, and violet wavelengths thatmeasured the various velocity components. Each wavelength wassplit into a pair, and one of each pair was frequency shifted toeliminate directional ambiguity of velocity measurements. Thebeams were then coupled into two fiber-optic probes that transmittedthem to the test section. The probes had both transmitting andreceiving optics and were operated in side-scatter mode, in which thereceiving optics from one probe were used to collect the scatteredlight from the opposite probe. The probes were fitted with beamexpanders and 750 mm lenses to allow the large required standoffdistance from the top to bottom end walls. The total angle betweenprobes was approximately 31–35 deg depending on measurement

location. The resulting ellipsoidal measurement volume at the beamcrossing had an estimated diameter D � 72 μm and a lengthL � 200 μm; at the highest shear locationmeasured, the diameter andlength in inner coordinates wasD� � 3 and L� � 8. Di-ethyl hexylsebacate was atomized to 1 μm diameter particles using a Laskinnozzle and injected upstream of the wind tunnel axial fan so that itwould be fully mixed into the flow at the measurement locations.At each measurement location, 20,000 coincident samples were

obtained. Instantaneous measurements were collected when theelapsed time between the center of one burst from a given channel tothe center of a burst from another channel was less than twice thesmallest burst duration recorded on any of the three channels. Thisresulted in an adaptive coincidence window that required that at aminimum a burst on one channel must started before the burst onanother channel ended; furthermore, it became more restrictive forhigh-velocity events (i.e., short burst durations) to improve thedetection of true coincidence. Comparisons of this method to a fixedcoincidence window of 10 μs at z� � 130 at location A1 in theboundary layer resulted in mean and rms values with a difference ofless than 1%. Sample rates were approximately 50 samples∕s nearthe wall and 200 samples∕s in the freestream. Instantaneousmeasurements were transformed from the nonorthogonal coordinatesystem of the probe setup to orthogonal cascade coordinates, andaverages of the results were corrected for velocity bias by using theinverse of the instantaneous velocitymagnitude as aweighting factor.Boundary-layer profiles were obtained at six locations across the

blade passage; three locations (A1–A3) at 0.2Cax downstream of thecascade leading edge plane, and three locations (B1–B3) at 1.03Cax.Figure 4 shows the locations overlaid on an oil streak-line pattern ofthe end wall flow from Lynch et al. [29]. The end wall flow patternshown in the figure displays the well-known secondary flow featuresdescribed by Langston et al. [19], including a saddle point at theseparation location of the inlet boundary layer, a strong passagevortex, and significant cross-stream velocity relative to the inviscidflow direction (which would follow the airfoil shape). Figure 4 also

Table 1 Inlet boundary-layerparameters

Variable Value at X∕Cax � −2.85δ99∕S 0.068θ∕S 0.0050Reθ 1550H 1.32Tu 5.7%Cf∕2 0.00229

Fig. 1 Closed-loop wind tunnel with corner turbine blade test section.

Table 2 Turbine blade geometry and operatingconditions

Condition Value

Scale 8.6XAxial chord Cax, m 0.218Span/axial chord S∕Cax 2.50Pitch/axial chord P∕Cax 0.826Inlet angle (βin, relative to axial direction), deg 35Exit angle (βexit, relative to axial direction), deg 60 degInlet Reynolds number (Rein � UinCax∕ν) 1.25 × 105

Inlet velocity Uin, m∕s 8.26Density ρ, incompressible, kg∕m3 1.16Kinematic viscosity, ν, m2∕s 1.56 × 10−5

Prandtl number Pr 0.71Exit Reynolds number (Reex � UexitCax∕ν) 2.00 × 105

Exit Mach number 0.04

956 LYNCH

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 4: Three-Dimensional Boundary Layer in a Turbine Blade Passage

indicates the cascade coordinate system, which will be used inpresenting the boundary-layer results. Note that in this coordinatesystemZ is thewall-normal direction for the boundary-layer profiles,and X and Y are parallel to the plane of the end wall.The oil film interferometry (OFI) method was used to directly

obtain wall shear stress at the six boundary-layer measurementlocations indicated in Fig. 4. The details of the OFI method aredescribed byLynch andThole [25]. In short, the thinning of a siliconeoil droplet as a function of time is directly related to the shear stressexerted by the airflow over the droplet. Bymeasuring the shear rate ofthe oil droplet, the wall shear at that location can be estimated. Thistechnique is extensively used in low- and high-speed wind tunnels(Naughton and Sheplak [34] and Driver [35]) and was benchmarkedbyLynch and Thole by comparing themeasured result to a Clauser fitto two-dimensional boundary-layer measurements from a LDV.Ruedi et al. [36] verified the technique in three-dimensional boundarylayers as well.Contours of the heat transfer coefficient over the entire end wall

surface were obtained on the end wall using the thin-film heatertechnique described by Lynch et al. [29]. A serpentine circuit heaterwith a thin copper top layer was attached to the end wall to supply a

uniform heat flux of 1000 W∕m2. The supplied heat flux was locallycorrected for conduction and radiation losses to estimate theconvective heat flux. The surface temperature of the heater wasmeasured with an infrared camera over multiple locations andcombined to create an overall surface temperature map. With theknown convective heat flux, surface temperature, and measuredfreestream temperature, the local heat transfer coefficient wasobtained and converted to nondimensional form. Note that the inletfreestream temperature was used as the reference temperature, sincethe cascade was low speed.Experimental uncertainties were estimated at a point in the end

wall boundary layer, using the partial derivative or propagation of theerror method for derived quantities (Moffat [37]). Table 3 lists theuncertainties for the various quantities reported in this paper. Notethat the percent uncertainty in mean W velocity is high because theabsolute value of that component was low.

IV. Description of Computations

The steady Reynolds-averaged Navier–Stokes (RANS) simu-lations by Lynch et al. [38] were interrogated at the boundary-layerlocations in Fig. 4 to assess predictive capability of the three-dimensional end wall boundary-layer velocity and turbulenceprofiles. In addition to the shear stress transport (SST) k − ωmodel ofMenter [26] used in Lynch et al. [38], a RSM (low-Reynolds stress-omega formulation for turbulence length scale) was considered forthis study, since it allows for anisotropy of the Reynolds stresses thatare important in this flowfield. The following describes theimplementation here (based on the formulation in ANSYS Fluent[39]); see the work by Lynch et al. [38] for additional details of theverification and validation.

3-axistraverse

Airfoil

Fig. 3 Three-component LDV mounted above the turbine cascade.

Cornervortex

SS HSseparation

PS HSseparation

A3

Y,V

X,UZ,W

A2

Passagevortex effect

A1

Saddlepoint

B3

B2

B 1

Fig. 4 Cascade boundary-layer measurement locations overlaid on oilflow streak lines (HS, horseshoe vortex).

-4

-3

-2

-1

0

1

0 0.2 0.4 0.6 0.8 1

Cp=

P s-P

s,in

/0.5

ρUin

2

s/stot

Prediction, SST k-ω

Blade 2

Blade 3

Blade 4

Blade 5

Blade 6

Fig. 2 Measured and predicted blade surface static pressure atmidspan.

Table 3 Typical measurement uncertainties

Variable Mean Total uncertainty Uncertainty, %

St 0.0043 0.00026 6.1Cf 0.00284 0.00023 8.1U 3.57 m∕s 0.041 m∕s 1.2V 1.70 m∕s 0.036 m∕s 2.1W −0.25 m∕s 0.043 m∕s 17.3u 0 1.17 m∕s 0.013 m∕s 1.2v 0 1.06 m∕s 0.013 m∕s 1.2w 0 0.73 m∕s 0.016 m∕s 2.2u 0v 0 −0.53 m2∕s2 0.0085 m∕s 1.6u 0w 0 −0.13 m2∕s2 0.0086 m∕s 6.4v 0w 0 0.10 m2∕s2 0.0063 m∕s 6.1

LYNCH 957

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 5: Three-Dimensional Boundary Layer in a Turbine Blade Passage

ANSYS Fluent, version 16, was used to solve the incompressibleNavier–Stokes equations for air using a steady RANS approach. Thesolver used a segregated pressure-based formulation, with theSIMPLE pressure-velocity coupling algorithm. Spatial gradientswere cell based using the least-squares option, and all solutionquantities were discretized using second-order upwinding. Airproperties were based on the experimental conditions of Lynch et al.[29], also indicated in Table 2. Because of only moderate changes inair temperature (<30°C difference between the end wall andmainstream) and the very low Mach number of the tunnel(Ma < 0.04), the properties were not assumed to vary, and viscousdissipation was not modeled. The steady incompressible Navier–Stokes equations in Reynolds-averaged form are given here inCartesian tensor format for brevity, where the index ranges from 1 to3 and 1 corresponds to the x direction, 2 corresponds to the ydirection, and 3 corresponds to the z direction in the results presentedlater. The modeled equations are

∂Ui

∂xi� 0 (1)

Uj

∂Ui

∂xj� −

1

ρ

∂P∂xi

� ∂∂xj

�ν

�∂Ui

∂xj� ∂Uj

∂xi

�− u 0

1u0j

�(2)

Uj

∂T∂xj

� ∂∂xj

�α∂T∂xj

− u 0jT

0�

(3)

The SST k − ωmodel byMenter [26], as implemented in ANSYSFluent, models the effect of the Reynolds stress and turbulent heatflux terms using the Boussinesq approximation:

−u 0i u

0j �

μtρ

�∂Ui

∂xj� ∂Uj

∂xi

�−2

3kδij (4)

−u 0jT

0 � μtρPrt

∂T∂xj

(5)

To determine the isotropic turbulent viscosity, the model solvesadditional equations for the turbulent kinetic energy k and specificdissipation rate ω:

Uj

∂k∂xj

� ∂∂xj

��ν� μt

ρσk

�∂k∂xj

�� �2μtSijSij� − ��β��fβ�kω� (6)

Uj

∂ω∂xj

� ∂∂xj

"�ν� μt

ρσω

�∂ω∂xj

#�

�α

μt�2μtSijSij�

�− �βiω2�

�"2�1 − F1�

1

ωσω;2

∂k∂xj

∂ω∂xj

#(7)

For Eq. (6), the terms in the square brackets on the right side are thediffusion, production, and dissipation of turbulent kinetic energy,respectively. Note that in the SST formulation of the k − ω model(Menter [26]), the piecewise function fβ� is defined to be a constantequal to 1. In Eq. (7), the square bracketed terms are the diffusion,production, dissipation, and the cross-diffusion (resulting from theblending of the k − ε and k − ωmodels; see thework ofMenter [26])of the specific dissipation rate, respectively. The turbulent viscosity isdefined as

μt �ρk

ω

1

max��1∕α��; �SF2∕a1ω��(8)

All model constants and blending functions in the previousequations (see the work of Menter [26,39]) are the standard valuesand formulations as implemented in Fluent.The RSM allows for nonisotropic eddy viscosity by solving

additional transport equations for each of the Reynolds stress tensorterms. In theory, this allows the model to better capture theanisotropic turbulence that occurs in the flowfield investigated here,and although it is more computationally expensive than steadyRANS, it is expected to provide more accurate predictions. Thetransport equation for the incompressible Reynolds stress terms in thelow-Reynolds stress-omega formulation is (neglecting buoyancy andsystem rotation effects)

Uk

∂�−u 0i u

0j�

∂xk� ∂

∂xk

��ν� μt

ρσk

� ∂�u 0i u

0j�

∂xk

−��−u 0

i u0k�∂Uj

∂xk� �u 0

ju0k�∂Ui

∂xk

��

�p 0�∂u 0

i

∂xj� ∂u 0

j

∂xi

��

−�2

3δijβ

�RSMkω

�(9)

where the bracketed terms on the right side of the equation are theviscous plus turbulent diffusion terms, the production term, thepressure-strain term, and the dissipation, respectively. Per the FluentTheory Guide [39], the pressure-strain term is modeled using a linearcombination of the contribution of slow and rapid pressure-straineffects, whilewall reflection effects are not necessary tomodel due tothe use of the specific dissipation and its properties close to a wall.The specific dissipation is modeled in the same way as described inEq. (7), and all model constants are the standard values given in [39].The model domain (see Fig. 5a) consisted of one passage of the

cascade from the end wall to half-span, with periodic boundaries onthe sides and a symmetry condition at half-span. To capture thedevelopment of the heat transfer coefficient on the end wall, the inletto the domain was located 4.3Cax upstream of the cascade leadingedge to extend upstream of the end wall heating that was supplied inthe experiment of Lynch et al. [29]. The boundary-layer codeTEXSTAN was used to generate profiles of mean velocity, turbulentkinetic energy, and turbulence dissipation (converted to specificdissipation) for the domain inlet, such that the calculated boundary-layer thickness at 2.85Cax upstream of the cascade matched the

Fig. 5 a) The computational domain for one periodic passage and b) view of the mesh.

958 LYNCH

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 6: Three-Dimensional Boundary Layer in a Turbine Blade Passage

measured result (Table 1). For the RSM,Reynolds stresses at the inletwere estimated from the specified turbulent kinetic energy assuminga two-dimensional turbulent boundary layer. The exit of the domainwas specified as an outflow boundary at 1.5Cax downstream of thetrailing-edge plane. The end wall and airfoil surfaces were specifiedas adiabatic no-slip walls, except a portion of the end wall startingfrom 3.4Cax upstream of the blade leading edge and ending at thedomain exit, which was given a constant wall heat flux of 1000 W∕m2

to match the same conditions established in the experiment of Lynchet al. [29].The computational grid (see Fig. 5b) was a multiblock hexahedral

body-fitted mesh created in ANSYS ICEM. Values of y� for all wallsurfaces were kept below 1, in accordance with the resolutionrequirements of the SST k − ωandRSMmodels. The inflation rate ofthe boundary layers was kept to 1.3 or less. The resulting model sizewas 1.2 million cells.Convergence of the simulations was determined by three metrics:

normalized residuals for the conservation equations had to reachvalues lower than 10−4 (10−6 for energy); the area-averaged Nusseltnumber on the end wall had to change less than 0.1% over 500iterations; and the mass-averaged exit total pressure downstream ofthe blade had to change less than 0.1% over 500 iterations. Asdescribed by Lynch et al. [38], a grid independence study wasperformed, and both the average end wall Nu and exit total pressurevaried by less than 1% for mesh cell counts of 0.62 million cells and3.1 million cells. The same grid independence was also verified forthe RSM model reported in this study. See the work by Lynch et al.[38] for validations of the simulation, including the midspan bladestatic pressure and the inlet end wall heat transfer coefficientdevelopment.

V. Results

Figure 6 compares contours of end wall heat transfer from theexperiment of Lynch et al. [29] to the predictions with the twoturbulence models. The locations of the end wall boundary-layerprofiles are also overlaid on the experimental heat transfer coefficientresults. All models agree well upstream of the blade passage asexpected, since the measured inlet boundary layer was used as aninput to the simulations. Note that the experiment of Lynch et al. [29]did not provide the components of the Reynolds shear stresses at theinlet, however; these were determined for the RSM model byassuming theywere similar to profiles in a two-dimensional flat plateboundary layer. The freestream turbulence in the experiment wasmoderately high (6% fromTable 1), whichwas not taken into accountin developing the inlet Reynolds stress profiles but was assumedto have a minor impact because of the long boundary-layerdevelopment length.

Further into the blade passage, however, there are significantdifferences in the model predictions relative to the experiment, withthe SST k − ωmodel generally predicting much higher heat transferlevels than seen in the experiment (also described by Lynch et al.[38]). The significant crossflow that exists in the blade passage causesthe sweeping of the heat transfer contours toward the suction side.Near the pressure side, the passage vortex that is developing causeshigh heat transfer near the blade pressure side junction. Both modelsalso predict a pressure-side separation near the leading edge, which isindicated by the island of low Nu extending out from the airfoilpressure surface near the leading edge. The RSMmodel does a betterjob than the SST k − ωmodel in capturing themodest increase in heattransfer toward the trailing edge of the passage.In addition to the spatially resolved heat transfer measurements,

discrete measurements of shear stress were obtained at the boundary-layer stations indicated in Fig. 4. Figure 7 summarizes the measuredand predicted friction coefficients and Stanton numbers at theboundary-layer stations. Note that the shear stress and heat transfercoefficient values in this figure are nondimensionalized to Cf and Stusing the local freestream velocity magnitude (determined fromboundary-layer profiles discussed later, see the nomenclature fordefinitions) and are written in a form that allows for comparison viathe modified Reynolds analogy (i.e., St�Pr2∕3 � Cf∕2). Figure 7indicates several interesting trends in the experiment. First, there areregions of the end wall near the pressure side (A1 and B1) where theReynolds analogy generally holds. The boundary layer near thepressure side of the airfoil is very new due to the crossflow in thepassage that progresses from the pressure to suction side and is not yetbeing strongly accelerated or skewed by the endwall secondary flow.However, for stations closer toward the suction side (A3 and B3),there are significant differences in the friction coefficient relative tothe heat transfer coefficient, due to strong pressure gradients in theblade passage that invalidate the use of the Reynolds analogy. Notethe significant variation in surface static pressure along the suctionside of the airfoil in Fig. 2. In particular, station A3 is close to thestrong favorable pressure that exists near the suction side leadingedge, and station B3 is close to the trailing-edge suction side, whichexperiences an adverse pressure gradient. It is well known (Kays andCrawford [40]) that the friction coefficient is more sensitive topressure gradient effects than the heat transfer coefficient, which isthe source of the discrepancy between Cf and St in Fig. 7 at stationsA3 and B3.The RSMmodel predictions in Fig. 7 agree towithin experimental

error for the friction coefficients at the forward stations in the bladepassage (A1–A3) and near the blade pressure side in the aft stations(B1), suggesting that this model can reasonably account for theeffects of the significant boundary-layer three dimensionality andpressure gradient effects at these locations. Lynch and Thole [25]

a)Experiment

(Lynch, et al. [29]) b) SST k-ω

(Lynch, et al. [38]) c) RSM

A1

A3 B1

B3

Fig. 6 a) Experimental measurements of end wall heat transfer; b) SST k − ω prediction; and c) RSM prediction, where St is normalized using thecascade inlet velocity.

LYNCH 959

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 7: Three-Dimensional Boundary Layer in a Turbine Blade Passage

indicated up to 70 deg of skew (defined as the angle between the endwall shear vector and the freestream mean flow vector) in theboundary layer at A1. The RSM model also accurately predicts thelocal value of St at A1 but slightly underpredicts St for stations A2,A3, and B1. However, the largest discrepancy betweenmeasured andpredicted friction coefficients is at stations B2 and B3. At theselocations, the crossflow is very significant (Fig. 4), and the flow is

being strongly accelerated through the throat of the airfoil passage.A later discussion will show how this impacts the velocity boundarylayer and how it is not properly modeled.Like the RSM model, the SST k − ωmodel does a reasonable job

predicting friction coefficients at stations A2 andA3 but significantlyoverpredicts Cf near the pressure side (A1 and B1). This may be inpart due to the anisotropy of the eddy viscosity in the highly skewed

Fig. 7 Comparison of measured and predicted heat transfer and shear stress (normalized using local boundary-layer edge velocity).

Fig. 8 Measured and predicted mean velocity components at stations a) A1 and b) A2.

Fig. 9 Normal Reynolds stresses for stations a) A1 and b) A2, and shear Reynolds stresses at stations c) A1 and d) A2.

960 LYNCH

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 8: Three-Dimensional Boundary Layer in a Turbine Blade Passage

boundary layer but could also be due to the lack of a transition modelin the SST k − ωmodel version used in this study. Several researchershave suggested that the flow near the blade pressure side in a turbineblade cascade is laminar (Harrison [21], Vera et al. [24], and Holleyet al. [41]). Toward the aft suction side (stations B2 and B3), the SSTk − ω model modestly overpredicts end wall heat transfer butseverely overpredicts shear stress, which, just like for the RSMmodel, is probably due to the difficulty of predicting the behavior ofthe highly accelerated 3D boundary layer at the passage exit.Velocity boundary-layer profiles were extracted from the

computational fluid dynamics (CFD) cases and compared toexperimental measurements, to explore reasons for the variousdisagreements between experiment and prediction in Fig. 7. Profilesof mean velocity at stations A1 and A2, expressed in innercoordinates, are shown in Fig. 8. Station A3 is not shown due to itssimilarity to A2. The profiles shown in these figures are presented inthe cascade coordinate system, whereX is the axial direction andZ isthewall-normal direction. In Fig. 8, the effect of boundary-layer threedimensionality is seen by comparing the U� and V� components,which are very different in the freestream but of similar magnitudetoward the end wall. This implies there is a significant amount ofcrossflow at station A1 (∼38 deg relative to the inviscid flowdirection). Both CFD model predictions do a reasonable jobcapturing mean velocity components, although the SST k − ωmodelunderpredicts the magnitude of the normalized U component atstation A1. The cause of this is due to the overprediction of the

friction coefficient for that model, which results in a large frictionvelocity, which is the normalizing variable in the inner coordinates.Reynolds stresses were also measured in the experiment and are

compared to the predictions for stationsA1 andA2 in Fig. 9.Note thatTable 3 indicates the measurement uncertainties, which areapproximately the same size as the markers in the figure. The RSMmodel provided the six components directly; for the SST k − ωmodel, they were estimated using the definition of the Reynoldsstress for isotropic eddy viscosity [see Eq. (4)], where the meangradients were obtained from the CFD solver directly. All Reynoldsstresses are normalized in inner coordinates in the figure. In general,both the normal and shear Reynolds stresses are larger at station A1relative to station A2, since station A1 is located near the origin of thepassage vortex, near the leading-edge blade pressure side. At A1, theRSMmodel does a better job predicting the peak in normal Reynoldsstresses at Z� ∼ 100, relative to the SST k − ω model. This is notunexpected, since the isotropic eddy viscosity assumption in the SSTk − ω model is not valid in this skewed boundary layer. However,neither model properly captures the high �u 0u 0� normal stresses, orthe cross-stream �u 0v 0� shear stresses, at either station A1 or A2. Thecross-stream �u 0v 0� fluctuations are a unique component of a three-dimensional boundary layer, which are particularly strong in aturbine blade cascade.The mean velocity boundary layer at the aft passage stations

(B1–B3) is difficult to measure due to its thinness (estimatedboundary-layer thickness of <2 mm, or 0.3% of the blade span).Figure 10 indicates the mean U- and V-velocity profiles, presentedhere in a coordinate system aligned with the local freestreamdirection (indicated by subscript FS on the velocity). The two-dimensional laminar and turbulent correlations are also included forreference. Because of the alignment with the local freestreamdirection, the V component goes to zero at the freestream, andnonzero V values within the boundary layer indicate threedimensionality (i.e., crossflow perpendicular to the freestream flowdirection). At station B1, there is only mild three dimensionality,which becomes more prevalent for B2 and B3. It is not entirely clearwhat state the boundary layer is in at these locations, but in thiscoordinate system, the streamwise velocity component (U�

FS) forstationB3 inFig. 10more closely resembles a laminar boundary layerthan a turbulent boundary layer. Langston et al. [19] also found thatthe cascade boundary layer near the exit of the blade passagewas verythin and conjectured that it was laminar, but they were also unable tofully sample the boundary layer with their hotwire system.Figure 11 compares the measured velocity profiles at stations B1

and B3 to the predictions from the two models in this study. Asexpected from the results in Fig. 7, there is good agreement for bothUand V components, among all models, for the velocity near the bladepressure side (B1). However, the effect of the boundary-layer threedimensionality and strong acceleration in the blade passage result inpoor predictions of the velocity profiles at station B3 near the suction

Fig. 10 Mean velocities at B1–B3, in a coordinate system aligned withthe freestream (Turb, turbulent).

Fig. 11 Predicted vs measured mean velocities for a) U�FS and b) V�

FS (Turb, turbulent).

LYNCH 961

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 9: Three-Dimensional Boundary Layer in a Turbine Blade Passage

side. In particular, the level of crossflow (V� component) is not wellcaptured. Furthermore, the SST k − ωmodel, used without a transitionmodel in the current study, does not reproduce the laminarlike behaviorof the streamwise velocity component (U�) at B3.

VI. Conclusions

The complex structure of a three-dimensional boundary layer in aturbine blade passage is challenging to capture with computationalsimulations. Predictions with a two-equation Reynolds-averagedNavier–Stokes (RANS) model and a full Reynolds-stress modelshow reasonable agreement in mean flow quantities particularly inthe forward part of the passage, but the predictions of the Reynoldsstress components are less satisfactory. Predictions are even lessreliable toward the aft section of the passage. Measurements indicatethat the cascade boundary layer there is likely laminar, which was notmodeled in the steady turbulent RANS formulations. Limited datasets exist for calibration of transition models in complex three-dimensional flows like the one investigated here, but future workshould include testing of those models to improve accuracy.Overall, there remain some deficiencies in the understanding of the

changes to near-wall turbulence with high levels of boundary-layerthree dimensionality, as well as the interaction of large-scale turbulentstructures generated at the end wall (passage secondary flows) with thethree-dimensional boundary layer. Predictions of the end wall heattransfer are still inadequate, given these limitations. It is hoped that thisdata set will provide motivation to developmodels that can account forsignificant boundary-layer three dimensionality.

Acknowledgments

The authorwould like to acknowledgeKarenThole andAtulKohlifor their technical input andwould like to thankUnitedTechnologies-Pratt & Whitney for permission to publish this work.

References

[1] Anderson, S. D., and Eaton, J. K., “Reynolds Stress Development inPressure-Driven Three-Dimensional Turbulent Boundary Layers,”Journal of Fluid Mechanics, Vol. 202, May 1989, pp. 263–294.doi:10.1017/S0022112089001187

[2] Coleman, G. N., Kim, J., and Spalart, P. R., “A Numerical Study ofStrained Three-Dimensional Wall-Bounded Turbulence,” Journal of

Fluid Mechanics, Vol. 416, Aug. 2000, pp. 75–116.doi:10.1017/S0022112000008806

[3] Moin, P., Shih, T.-H., Driver, D., andMansour, N.N., “DirectNumericalSimulation of a Three-Dimensional Turbulent Boundary Layer,”Physics of Fluids A: Fluid Dynamics, Vol. 2, No. 10, 1990,pp. 1846–1853.doi:10.1063/1.857658

[4] Schwarz,W. R., andBradshaw, P., “Turbulence Structural Changes for aThree-Dimensional Turbulent Boundary Layer in a 30° Bend,” Journalof Fluid Mechanics, Vol. 272, Aug. 1994, pp. 183–210.doi:10.1017/S002211209400443X

[5] Flack, K. A., and Johnston, J. P., “Near-Wall Flow in a Three-Dimensional Boundary Layer on the Endwall of a 30° Bend,”Experiments in Fluids, Vol. 24, No. 2, 1998, pp. 175–184.doi:10.1007/s003480050164

[6] Johnston, J. P., and Flack, K. A., “Review—Advances in Three-Dimensional Turbulent Boundary Layers with Emphasis on the Wall-Layer Regions,” Journal of Fluids Engineering, Vol. 118, No. 2, 1996,pp. 219–232.doi:10.1115/1.2817367

[7] Flack, K. A., “Near-Wall Structure of Three-Dimensional TurbulentBoundary Layers,” Experiments in Fluids, Vol. 23, No. 4, 1997,pp. 335–340.doi:10.1007/s003480050119

[8] Bruns, J. M., Fernholz, H. H., and Monkewitz, P. A., “An ExperimentalInvestigation of a Three-Dimensional Turbulent Boundary Layer in an‘S’-Shaped Duct,” Journal of Fluid Mechanics, Vol. 393, 1999,pp. 175–213.doi:10.1017/S0022112099005522

[9] Olcmen,M. S., and Simpson, R. L., “AnExperimental Study of a Three-Dimensional Pressure-Driven Turbulent Boundary Layer,” Journal of

Fluid Mechanics, Vol. 290, 1995, pp. 225–262.doi:10.1017/S0022112095002497

[10] Olcmen, M. S., and Simpson, R. L., “Perspective: On the Near WallSimilarity of Three-Dimensional Turbulent Boundary Layers (DataBank Contribution),” Journal of Fluids Engineering, Vol. 114, No. 4,1992, pp. 487–495.doi:10.1115/1.2910059

[11] Abrahamson, S. D., and Eaton, J. K., “Heat Transfer Through aPressure-Driven Three-Dimensional Boundary Layer,” Journal of HeatTransfer, Vol. 113, No. 2, 1991, pp. 355–362.doi:10.1115/1.2910569

[12] Lewis,D. J., and Simpson, R. L., “Turbulence Structure ofHeat TransferThrough a Three-Dimensional Turbulent Boundary Layer,” Journal ofThermophysics and Heat Transfer, Vol. 12, No. 2, 1998, pp. 248–255.doi:10.2514/2.6328

[13] Eibeck, P. A., and Eaton, J. K., “Heat Transfer Effects of a LongitudinalVortex Embedded in a Turbulent Boundary Layer,” Journal of Heat

Transfer, Vol. 109, No. 1, 1987, pp. 16–24.doi:10.1115/1.3248039

[14] Wroblewski, D. E., and Eibeck, P. A., “Measurements of Turbulent HeatTransport in a Boundary Layer with an Embedded Streamwise Vortex,”International Journal of Heat and Mass Transfer, Vol. 34, No. 7, 1991,pp. 1617–1631.doi:10.1016/0017-9310(91)90141-Z

[15] Shizawa, T., and Eaton, J. K., “Turbulence Measurements for aLongitudinal Vortex Interacting with a Three-Dimensional TurbulentBoundary Layer,” AIAA Journal, Vol. 30, No. 1, 1992, pp. 49–55.doi:10.2514/3.10881

[16] Langston, L. S., “Crossflows in a Turbine Cascade Passage,” Journal ofEngineering for Power, Vol. 102, No. 4, 1980, pp. 866–874.doi:10.1115/1.3230352

[17] Sharma, O. P., and Butler, T. L., “Predictions of Endwall Losses andSecondary Flows in Axial Flow Turbine Cascades,” Journal of

Turbomachinery, Vol. 109, No. 2, 1987, pp. 229–236.doi:10.1115/1.3262089

[18] Wang, H. P., Olson, S. J., Goldstein, R. J., and Eckert, E. R. G., “FlowVisualization in a Linear Turbine Cascade of High Performance TurbineBlades,” Journal of Turbomachinery, Vol. 119, No. 1, 1997, pp. 1–8.doi:10.1115/1.2841006

[19] Langston, L. S., Nice, M. L., and Hooper, R. M., “Three-DimensionalFlow Within a Turbine Cascade Passage,” Journal of Engineering for

Power, Vol. 99, No. 1, 1977, pp. 21–28.doi:10.1115/1.3446247

[20] Kang,M.B., andThole,K.A., “FlowfieldMeasurements in theEndwallRegion of a Stator Vane,” Journal of Turbomachinery, Vol. 122, No. 3,2000, pp. 458–466.doi:10.1115/1.1303703

[21] Harrison, S., “Secondary Loss Generation in a Linear Cascade of High-Turning Turbine Blades,” Journal of Turbomachinery, Vol. 112, No. 4,1990, pp. 618–624.doi:10.1115/1.2927702

[22] Gregory-Smith, D. G., Graves, C. P., and Walsh, J. A., “Growth ofSecondary Losses and Vorticity in an Axial Turbine Cascade,” Journalof Turbomachinery, Vol. 110, No. 1, 1988, pp. 1–8.doi:10.1115/1.3262163

[23] Knezevici, D. C., Sjolander, S. A., Praisner, T. J., Allen-Bradley, E., andGrover, E. A., “Measurements of Secondary Losses in a TurbineCascade with the Implementation of Nonaxisymmetric EndwallContouring,” Journal of Turbomachinery, Vol. 132, No. 1, 2010,Paper 011013-1-10.doi:10.1115/1.3072520

[24] Vera, M., Blanco, E. D. L. R., Hodson, H., and Vazquez, R., “EndwallBoundary Layer Development in an Engine Representative Four-StageLow Pressure Turbine Rig,” Journal of Turbomachinery, Vol. 131,No. 1, 2009, Paper 011017.doi:10.1115/1.2952382

[25] Lynch, S. P., and Thole, K. A., “Comparison of the Three-DimensionalBoundary Layer on Flat Versus Contoured Turbine Endwalls,” Journalof Turbomachinery, Vol. 138, No. 4, 2016, Paper 041008-1-10.doi:10.1115/1.4032165

[26] Menter, F. R., “Two-Equation Eddy-Viscosity Turbulence Models forEngineering Applications,” AIAA Journal, Vol. 32, No. 8, 1994,pp. 1598–1605.doi:10.2514/3.12149

[27] Wilcox, D. C., Turbulence Modeling for CFD, DCW Industries, LaCanada, CA, 1998.

[28] Lynch, S. P., “The Effect of Endwall Contouring on Boundary LayerDevelopment in a Turbine Blade Passage,” Ph.D. Thesis, VirginiaPolytechnic Inst. and State Univ., Blacksburg, VA, 2011.

[29] Lynch, S. P., Sundaram, N., Thole, K. A., Kohli, A., and Lehane, C.,“Heat Transfer for a Turbine Blade with Nonaxisymmetric Endwall

962 LYNCH

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2

Page 10: Three-Dimensional Boundary Layer in a Turbine Blade Passage

Contouring,” Journal of Turbomachinery, Vol. 133, No. 1, 2011,Paper 011019.doi:10.1115/1.4000542

[30] Bons, J. P., Sondergaard, R., and Rivir, R. B., “Turbine SeparationControl Using Pulsed Vortex Generator Jets,” Journal of Turbomachi-nery, Vol. 123, No. 2, 2001, pp. 198–206.doi:10.1115/1.1350410

[31] McAuliffe, B. R., and Sjolander, S. A., “Active Flow Control UsingSteady Blowing for a Low-Pressure Turbine Cascade,” Journal of

Turbomachinery, Vol. 126, No. 4, 2004, pp. 560–569.doi:10.1115/1.1791291

[32] Praisner, T. J., Allen-Bradley, E., Grover, E. A., Knezevici, D. C., andSjolander, S.A., “Application ofNonaxisymmetric Endwall Contouringto Conventional and High-Lift Turbine Airfoils,” Journal of

Turbomachinery, Vol. 135, No. 6, 2013, Paper 061006.doi:10.1115/1.4024023

[33] Mensch, A., and Thole, K. A., “Overall Effectiveness of a Blade Endwallwith Jet Impingement and Film Cooling,” Journal of Engineering for GasTurbines and Power, Vol. 136, No. 3, 2013, Paper 031901.doi:10.1115/1.4025835

[34] Naughton, J. W., and Sheplak, M., “Modern Developments inShear-Stress Measurement,” Progress in Aerospace Sciences, Vol. 38,Nos. 6–7, 2002, pp. 515–570.doi:10.1016/S0376-0421(02)00031-3

[35] Driver, D. M., “Application of Oil-Film Interferometry Skin-FrictionMeasurement to Large Wind Tunnels,” Experiments in Fluids, Vol. 34,

No. 6, 2003, pp. 717–725.doi:10.1007/s00348-003-0613-1

[36] Ruedi, J. D.,Nagib,H.,Österlund, J., andMonkewitz, P.A., “Evaluationof Three Techniques for Wall-Shear Measurements in Three-Dimensional Flows,” Experiments in Fluids, Vol. 35, No. 5, 2003,pp. 389–396.doi:10.1007/s00348-003-0650-9

[37] Moffat, R. J., “Describing the Uncertainties in ExperimentalResults,” Experimental Thermal and Fluid Science, Vol. 1, No. 1,1988, pp. 3–17.doi:10.1016/0894-1777(88)90043-X

[38] Lynch, S. P., Thole, K. A., Kohli, A., and Lehane, C., “ComputationalPredictions of Heat Transfer and Film-Cooling for a Turbine BladewithNonaxisymmetric Endwall Contouring,” Journal of Turbomachinery,Vol. 133, No. 4, 2011, Paper 041003.doi:10.1115/1.4002951

[39] Theory Guide, ANSYS Fluent, Ver. 16.1.[40] Kays,W.M., andCrawford,M. E.,ConvectiveHeat andMass Transfer,

McGraw–Hill, New York, 1980.[41] Holley, B. M., Becz, S., and Langston, L. S., “Measurement and

Calculation of Turbine Cascade Endwall Pressure and Shear Stress,”Journal of Turbomachinery, Vol. 128, No. 2, 2006, pp. 232–239.doi:10.1115/1.2137744

N. L. KeyAssociate Editor

LYNCH 963

Dow

nloa

ded

by P

EN

NSY

LV

AN

IA S

TA

TE

UN

IVE

RSI

TY

on

Apr

il 13

, 201

8 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.B

3623

2