2017 - Endwall Loss in Turbine Cascade

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Endwall Loss in Turbine

Cascades
Prior to the detailed design of components, turbomachinery engineers must guide a
mean-line or throughflow design toward an optimum configuration. This process requires
a combination of informed judgement and low-order correlations for the principle sour-
John D. Coull ces of loss. With these requirements in mind, this paper examines the impact of key design
Whittle Laboratory, parameters on endwall loss in turbines, a problem which remains poorly understood.
University of Cambridge, This paper presents a parametric study of linear cascades, which represent a simplified
Cambridge CB3 0DY, UK model of real-engine flow. The designs are nominally representative of the low-pressure
e-mail: jdc38@cam.ac.uk turbine blades of an aero-engine, with varying flow angles, blade thickness, and suction
surface lift styles. Reynolds-averaged Navier–Stokes (RANS) calculations are performed
for a single aspect ratio (AR) and constant inlet boundary layer thickness. To character-
ize the cascades studied, the two-dimensional design space is examined before studying
endwall losses in detail. It is demonstrated that endwall loss can be decomposed into two
components: one due to the dissipation associated with the endwall boundary layer and
another induced by the secondary flows. This secondary-flow-induced loss is found to
scale with a measure of streamwise vorticity predicted by classical secondary flow
theory. [DOI: 10.1115/1.4035663]

1 Introduction generated where the cross-flow separates at the blade–endwall junc-


tions. Downstream of the trailing edge (TE), the vortex system inter-
Endwall effects generate around a third of the total aerody-
acts with the trailing shed vorticity, which arises due to the variation
namic loss in a turbine [1]. These losses are generated by dissipa-
in loading along the blade span (e.g., Ref. [7]). Deviations to this
tion in the hub and casing boundary layers and by the secondary
basic flow pattern have been reported by various authors, particu-
flow, which is generated as the nonuniform inlet flow turns
larly in regards to the interaction of the two main vortex legs and
through the blade row.
the smaller flow features, suggesting that the detailed behavior is
Much of the fundamental understanding of endwall flows has
design-dependent. Furthermore, this description of the mean flow
been built on studies of linear cascades, which represent an approxi-
field neglects any unsteady effects that will be present.
mate model for real-engine behavior. Early theoretical work focused
At the preliminary design stage, the losses introduced by the
on predicting the secondary flows in an inviscid flowfield by consid-
endwall flow must be modeled, prior to detailed design of the
ering the convection of vortex filaments through a cascade (e.g.,
blade shape. This process must therefore rely on low-order loss
Ref. [2]). More detailed understanding of the flow behavior was
correlations. Several such methods are available in the open litera-
gleaned over many decades of experimental research, as summar-
ture, for example, Ainley and Mathieson [8], Dunham and Came
ized by Sieverding [3] and Langston [4].
[9], Kacker and Okapuu [10], Craig and Cox [11], and Traupel
Figure 1 shows a diagram based on one of the current RANS
[12]. Many of these correlations remain in use today, but they are
simulations, where vortex structures have been visualized using
the eigenvalue method of Adrian et al. [5]. The endwall is subject
to the same static pressure field as the freestream flow; the incom-
ing endwall boundary layer therefore separates in the adverse
pressure gradient approaching the blade leading edge and rolls up
into a horseshoe vortex (HSV). Once formed, the trajectory of this
vortex largely follows the static pressure gradients: the suction-
side leg (HSV-SS) initially remains close to the blade suction sur-
face corner, while the pressure-side leg (HSV-PS) is driven across
the passage toward the low-pressure region on the adjacent blade
suction surface. This leg entrains fluid from the fresh endwall
boundary layer downstream of the HSV separation and forms into
the main passage vortex (PV).
On reaching the blade suction surface, the passage vortex inter-
acts with the HSV-SS (counter vortex (CV)) from the adjacent
horseshoe vortex and the blade surface, inducing large velocity gra-
dients and high dissipation rates in this region [6]. The two vortices
then move up the blade suction surface toward midspan. Other
smaller vortical structures are also observed: additional vortices are
induced on the blade suction surface by the interaction of the pas-
sage and counter vortices; the HSV is actually a system of counter-
rotating vortices, as is evident in Fig. 1; and corner vortices are often

Contributed by the International Gas Turbine Institute (IGTI) of ASME for


publication in the JOURNAL OF TURBOMACHINERY. Manuscript received October 24,
2016; final manuscript received December 16, 2016; published online March 15,
2017. Editor: Kenneth Hall. Fig. 1 Secondary flow in a turbine cascade

Journal of Turbomachinery Copyright V


C 2017 by ASME AUGUST 2017, Vol. 139 / 081004-1

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


Fig. 3 Mises Cp distributions for the designs in Fig. 2: (a) rela-
tive to outlet pressure and (b) relative to trailing edge pressure

With no consistent trends available from modeling or published


data, the low-order design sensitivities for turbine endwall loss
remain poorly understood. The author believes that the best way
to address this problem is through the study of parametric designs.
A turbine cascade has at least 13 key independent design and oper-
ating parameters (see Table 1), which set the flow angles, aspect
ratio, operating condition, suction surface pressure distribution,
and blade thickness. Exploring this multidimensional design space
Fig. 2 Designs with varying flow angles, see Table 1 experimentally, or with high-order techniques such as large eddy
simulation, would be a huge undertaking and is unlikely to be via-
based on the circular-arc blading typical of pre-1970 designs and ble. Instead this paper uses an automated RANS approach to study
some have little physical basis. More recently, Benner et al. [13] a large number of designs. Nonetheless, it is infeasible to cover all
formulated a correlation based on a survey of experimental cas- of the design parameters in a single paper. A single aspect ratio is
cade data. However, the dataset was relatively sparse and did not studied, as it is known that endwall loss is inversely proportional
exhibit clear trends; furthermore, there were several inconsisten- to the aspect ratio for sufficiently large values (e.g., see Ref. [15]).
cies in the measurement and processing techniques for each cas- A turbulent collinear boundary layer of fixed height is set at the
cade. Thus, it was not clear which design parameters needed to be inlet for all the cases; therefore, this paper does not consider the
included, or what form the correlation should take. Coull and impact of inlet conditions, which can impact the endwall flow in
Hodson [14] showed that the available correlations vary widely in cascades and engines (e.g., see Ref. [6]).
their predictions and their sensitivity to key design parameters. Section 2 details the design and CFD methods employed. The
subsequent sections focus on the impact of flow angles, examining

Table 1 Key design parameters for the blades in Fig. 2, with approximate tolerances used in the Mises design code

Target Tolerance

Inlet flow angle, a1 20 deg ! 40 deg —


Outlet flow angle, a2 50 deg ! 70 deg 60.02 deg
Midspan exit Mach number, M2 0.7 60.005
Reynolds number, ReCx 200,000 61%
Aspect ratio, h=Cx 3 —
Inlet momentum thickness, h=Cx 0.01 —
Suction side, Cp Diffusion factor, DF 0.28 60.001
Peak suction location,pPSL
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 0.52 60.005
Leading edge ratio, CPLE =Cppeak 0.7a —
Thickness Max. thick., Tmax =Cx 0.20 60.02
Max. thick. location (x=Cx ) 0.35 60.04
TE thickness, t=Cx 0.02 0
TE wedge angle 9 deg 60.2 deg
a
This parameter varies to achieve incidence tolerance (Fig. 3).

081004-2 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


the two-dimensional design space (Sec. 3) and profile losses input parameters to maximize a mesh quality factor, which incor-
(Sec. 4) in order to fully characterize the cascades studied. End- porates measures of cell skew, aspect ratio, and expansion ratio.
wall loss is then discussed in Sec. 5 for a larger range of designs. Steady RANS calculations are performed with the Rolls-Royce
in-house solver HYDRA; full details of this code are given by
Moinier and Giles [23]. The spatial discretization is based on an
2 Numerical Methods upwind edge-based finite volume scheme and is second-order
accurate. Calculations are performed using the two-equation
2.1 Parametric Blade Design. An iterative two-dimensional shear-stress-transport turbulence model. For simplicity, calcula-
blade design code has been developed which performs simulations tions are performed with fully turbulent boundary layers. While
using the coupled Euler and boundary layer integral solver, Mises closer agreement with experimental data can be obtained using a
[16,17]. The code is designed to carefully control the freestream transition model [24], the fully turbulent calculations give similar
velocity distribution over the suction surface boundary layer, endwall loss predictions and are more robust.
which is the key for controlling the profile loss (e.g., Ref. [18]). Calculations are performed on a half-passage domain for each
The code adjusts the blade shape and pitch to achieve the desired linear cascade, with an inviscid wall at midspan to provide sym-
suction surface aerodynamics and exit flow conditions. The thick- metry. Figure 4 shows the domain: the inlet is located at a distance
ness distribution of the blade is also specified. To ensure the of 0:8Cx upstream of the leading edge and the domain outlet is
designs are tolerant to incidence, additional Mises calculations 1:2Cx downstream of the trailing edge. Spanwise distributions of
with 10 deg of positive incidence are executed. Leading edge total pressure, total temperature, flow angles, and turbulence
overspeed spikes are minimized by using a continuous curvature parameters are applied at the inlet and static pressure at the exit. A
geometry [19] and by iteratively drooping the blade leading edge. collinear turbulent boundary layer is specified at the inlet with
Figure 2 shows a set of 35 blades produced by varying inlet and momentum thickness equal to 0:01Cx ; the distributions are
exit flow angles while keeping all other input parameters constant. obtained from a separate flat-plate boundary layer calculation.
The key design parameters for this set are shown in Table 1. The The freestream flow turbulence parameters are set to obtain a tur-
suction surface aerodynamic parameters (DF ¼ 0.28 and bulence level of 5% with an integral length scale of 0:1Cx , which
PSL ¼ 0.52) correspond to design D studied by Coull et al. [18], is broadly representative of a multistage turbine [25].
which is nominally representative of modern low-pressure turbine The current calculations do not capture the impact of flow
designs. The surface pressure coefficient (Cp ) distributions pre- unsteadiness on the loss generation, which will introduce some
dicted by Mises are shown in Fig. 3(a); boundary layer transition uncertainty in the results. Denton and Pullan [6] examined a linear
on the suction surface has been specified at the peak velocity loca- stage and found that the loss development in the first blade row
tion (52%). It is clear from this graph that the trailing edge pres- was approximately equal in steady and unsteady RANS calcula-
sure varies significantly around the design space, a factor that has tions. Therefore, the contribution of the unsteadiness to cascade
previously been observed by other authors (e.g., Ref. [20]). Nor- endwall loss is believed to be small; in general, this will not be
malizing Cp by the trailing edge value, Fig. 3(b) demonstrates that the case in a real turbine where additional unsteadiness is induced
the suction surface CpTE distribution is approximately constant by the interaction of adjacent blade rows.
for all of the designs. Some differences are noted in the leading
edge region of the suction surface for the three designs in the top- 2.3 Loss Coefficients. Performance statistics are calculated
right corner of Fig. 2. These designs have the lowest pitch, and using the “outlet” plane shown in Fig. 4, which is located 0:5Cx
therefore the blade thickness introduces significant blockage, behind the trailing edge. A constant-area mixing calculation is
increasing the velocity in the early portion of the passage above
performed. The entropy loss coefficient [1] is defined as the net
the aerodynamic target. Significant variation in pressure surface increase between the mixed-out entropy at inlet (s1;mix ) and exit
loading is observed in Fig. 3 and will be discussed in Sec. 3.2. (s2;mix ):
The fundamental Reynolds number for profile loss is that based
on the suction surface length (ReS0 ), which characterizes the suc- nnet ¼ ðs2;mix  s1;mix Þ ½T2;mix =ðh02;mix  h2;mix Þmid (1)
tion surface boundary layer. This length is not known a priori until
a profile is designed; in contrast, the axial chord is selected much where h is the specific enthalpy. This formulation aligns with the
earlier in the design process. For high aspect ratio blades, the end- findings of Sharma and Butler [15], who demonstrated for several
wall loss is largely insensitive to Reynolds number, e.g., Ref. cascades that the net mass-averaged loss was approximately inde-
[21]; therefore, for simplicity, the calculations are performed at pendent of the inlet boundary layer thickness. Note that the mid-
constant ReCx . span exit conditions are used as a reference in Eq. (1), rather than
the passage-averaged conditions. This approach has the advantage
2.2 Three-Dimensional Meshing and CFD. Automated of ensuring that the profile loss is consistently defined between
meshing is performed using an optimizer built around the Rolls- Mises and RANS calculations.
Royce PADRAM code [22], which combines a blade O-mesh The total passage loss is split into profile (midspan) and endwall
with multi-block passage H-meshes. The optimizer drives the loss using the traditional formulation:

nendwall ¼ nnet  nprofile (2)

where nprofile is calculated from the midspan inlet and exit condi-
tions using Eq. (1). This breakdown is somewhat arbitrary and can
produce strange results, such as negative endwall loss for low
aspect ratios (e.g., Ref. [26]) or thick inlet boundary layers. The
relatively high aspect ratio (h=Cx ¼ 3) and thin inlet boundary
layer (h=Cx ¼ 0.01) used in this paper avoid such problems, none-
theless it is recognized that the breakdown of profile and endwall
loss is problematic.

2.4 Mesh Sensitivity and Validation. Figure 5 shows the


calculated loss coefficients as the mesh is refined for a single
Fig. 4 Typical CFD domain blade design. The viscous O-mesh and radial distributions are set

Journal of Turbomachinery AUGUST 2017, Vol. 139 / 081004-3

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


50 deg (1.3) and a2 ¼ 70 deg (2.0). This length is crucial
for the profile loss since it is the characteristic length scale of the
suction surface boundary layer, which dominates the loss produc-
tion. Furthermore, since ReCx is held constant, this ratio sets the
characteristic Reynolds number ReS0 , which varies between
260,000 (a2 ¼ 50 deg) and 400,000 (a2 ¼ 70 deg).

3.2 Pitch and Loading Coefficients. The spacing of each


blade is determined in the design process by the aerodynamic and
thickness targets (see Sec. 2.1) and varies with the flow angles.
Figure 7(b) shows the ratio of pitch to axial chord (s=Cx ) which
varies by a factor of around 2, increasing for designs with higher
exit angle and lower turning. Figure 7(c) shows the ratio of pitch
Fig. 5 Mesh sensitivity for a single design (a1 5 30 deg, to suction surface length (s=S0SS ), which varies less strongly and
a2 5 265 deg, and Tmax =Cx 5 0.15) is largely dependent on the inlet flow angle. Circulation arguments
show that this ratio is closely related to the turning [14].
The blade pitch may be considered in terms of a lift coefficient.
to maintain a maximum yþ value of approximately unity on the Zweifel [29] calculated the ratio of the blade tangential force to
blade and endwall surfaces, with an expansion ratio of 1.14. The an ideal case (PSS ¼ P2 and PPS ¼ P01 ):
final template has 10.2 million cells for this case, and the passage,
profile, and endwall loss coefficients are within 0.00005 of the fin- þ    
x s q2 V22 cos a2  V1 
 (3)
est mesh values. Zw ¼ Cp d ¼  sin a 2  sin a 1
To assess the accuracy of the CFD setup, calculations are Cx Cx ðP01  P2 Þ V2
performed for the low-pressure turbine cascade of Hodson and
Dominy [27] and the high-pressure turbine design studied by As shown in Fig. 8(a), this parameter varies by around 40–45%
Gregory-Smith et al. [28]. Figure 6 compares the measured and across the design space, being higher for designs with larger a1 .
predicted endwall losses for different cases; the average error is The variation can be understood by splitting the Zweifel coeffi-
less than 11%, which is deemed to be acceptable. cient into the negative contribution of the pressure surface integral
(Figs. 8(b) and 8(c)) and the positive contribution of the suction
surface integral (Figs. 8(d) and 8(e)):
3 The Two-Dimensional Design Space
The designs in Fig. 2 represent a family of blades with similar "ð  # "ð  #
suction-side aerodynamics. This section examines the variation x x
Zw ¼ Cp d  Cp d (4)
within the two-dimensional design space, prior to the discussion SS Cx PS Cx
of profile (Sec. 4) and endwall losses (Sec. 5).
The pressure-side integral (Fig. 8(b)) is higher for low-turning
3.1 Surface Length. Figure 7(a) shows the ratio of suction designs, in line with the distributions in Fig. 3. The Appendix
surface length to axial chord (S0SS =Cx ), which is largely a func- presents a simple model that demonstrates that this variation is
tion of the exit flow angle, increasing rapidly between a2 ¼ driven by the cross-passage pressure gradient in the covered turn-
ing region: in general, the pressure surface velocity will be higher
for designs with less deflection and greater thickness. The suction-
side integral in Fig. 8(d) demonstrates a variation of around 40%
across the design space, despite the similarity of the suction-side
Cp distributions in Fig. 3. This apparent contradiction arises
because the integral in Eq. (4) is performed with respect to the
axial direction; as shown in Fig. 9, the distributions of Cp with
axial distance vary significantly with flow angles.
Coull and Hodson [14] argued that the Zweifel coefficient did
not adequately account for airfoil camber and proposed an alterna-
tive circulation coefficient:

Fig. 6 CFD and experimental endwall losses [27,28]

Fig. 7 Geometric parameters for the designs in Fig. 2: (a) suc-


tion surface length versus axial chord, (b) pitch versus axial Fig. 8 (a) Zweifel lift coefficient [29] and breakdown (a) Zweifel
chord, (c) pitch versus suction surface length, and (d) suction lift coefficient [29] and breakdown: pressure-side integral ((b),
surface length versus effective throat (c)) and suction-side integral ((d), (e))

081004-4 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


Sec. 3.4). From this comparison, one may therefore conclude that
blade loading coefficient and pitch are determined by the specified
suction side Cp design and the pressure-side loading, which is pre-
dominantly a function of flow angles and thickness (Appendix).

3.3 Incidence and Deviation. As discussed in Sec. 2.1, the


design code iteratively droops the leading edge of the blade to pre-
vent overspeed spikes at design and þ10 deg flow incidence. The
resultant leading edge incidence, defined as the difference
between the inlet flow and metal angles, is presented in Fig. 11(a).
This parameter is small for much of the design space, but rises
significantly for the highest exit angle designs. It should be noted
that a manual design of the leading edge shape may produce dif-
ferent results. The deviation relative to the average exit metal
angle is presented in Fig. 11(b) and remains small across the
design space.

3.4 Trailing Edge Pressure and Mach Number. At the


trailing edge plane of a blade row, the freestream velocity tends to
be higher than the downstream value because of the combined
blockage of the boundary layers and the trailing edge [1]. How-
ever, the pressure across this plane is usually nonuniform due to
the potential field of the blades, and several authors report high
static pressure in the locality of the trailing edge (e.g., Ref. [20]).
This effect has two competing implications for the blade profile
loss. First, for a given peak Mach number, the suction surface
Fig. 9 Mises Cp versus x =Cx for the designs in Fig. 2: (a) rela-
boundary layer must undergo additional deceleration, increasing
tive to outlet pressure and (b) relative to trailing edge pressure
the diffusion factor (DF), trailing edge momentum thickness [30],
and hence the loss. This effect is partly offset by the lower free-
þ  
pffiffiffiffiffiffi S stream velocity at the boundary layer edge, which implies a lower
Co ¼ Cp d (5) absolute momentum deficit. Second, higher static pressure around
S0SS
the trailing edge implies higher base pressure on the trailing edge
As shown in Fig. 10(a), this parameter exhibits a different varia- radius, thus the base pressure loss can be reduced and may be neg-
tion across the design space compared to Zweifel (Fig. 8(a)). The ative, e.g., Refs. [31] and [32].
circulation coefficient can also be decomposed to separate the The variation in trailing edge Mach number is shown in Fig.
pressure and suction side contributions: 12(a) for the Mises calculations; the RANS calculations give a
very similar result. It can be observed that the Mach number
decreases as the exit flow angle increases in magnitude. To a first
ð     ð   order, this effect is driven by the uncovered turning angle down-
pffiffiffiffiffiffi S S0PS pffiffiffiffiffiffi S
Co ¼ Cp d  Cp d stream of the blade throat, defined in Fig. 13 and plotted in
SS S0SS S0SS PS S0PS Fig. 12(b), which necessitates a pressure gradient normal to the
(6) uncovered portion of the suction surface. Figure 12(c) shows the
strong correlation between the uncovered turning and the trailing
The ratio of pressure to suction side lengths S0PS =S0SS is rela- edge Mach number for this set of designs. This relationship is not
tively constant across the design space (0.86 6 0.06). The universal and depends on the blade thickness and suction surface
pressure-side integral (Figs. 10(b) and 10(c)) shows a similar trend Cp distribution.
to Fig. 8(b), but the variation is larger because of the square root
in Eq. (6). As would be expected from the data in Fig. 3, the
suction-side integral is relatively constant, varying by about 10% 4 Profile Loss
in line with the local Mach number at the trailing edge (see Figures 14(a) and 14(b) present the mixed-out profile losses
calculated using RANS and Mises, respectively, both with fully
turbulent boundary layers. While there are differences in the level
of loss, the trends of the two predictions agree: in general, the

Fig. 10 (a) Circulation lift coefficient [14] and breakdown: pres- Fig. 11 (a) Leading edge incidence (difference between the
sure-side integral ((b), (c)) and suction-side integral ((d), (e)) inlet flow and metal angles) and (b) deviation (Mises)

Journal of Turbomachinery AUGUST 2017, Vol. 139 / 081004-5

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


Fig. 15 Breakdown of loss contributions in Eq. (7) for fully tur-
Fig. 12 (a) Trailing edge Mach number (Mises); (b) uncovered bulent Mises calculations: (a) momentum deficit; (b) blockage;
turning angle; (c) codependence (c) base pressure

The terms in Eq. (7) have been extracted from the fully turbu-
lent Mises calculations and are shown in Fig. 15. The first term
(Fig. 15(a)) is the largest and represents the boundary layer
momentum deficit. This term is dominated by the suction-side
contribution which, for a fully turbulent boundary layer, scales as

2 hSS 2 S0SS
Fig. 13 Definition of uncovered turning angle ¼ f ðCpSS Þ Re0:2
S0 (8)
s cos b2 s cos b2

where f ðCpSS Þ will be a constant for designs with matching Cp


distributions. The variation in Fig. 15(a) is therefore largely
driven by the geometric ratio of suction surface length to the
effective throat (S0SS =s cos b2 ), presented in Fig. 7(d). Smaller
effects are introduced by the variation in Reynolds number and
pressure surface velocity (e.g., see Fig. 10(b)).
The base pressure loss in Fig. 15(c) is the smallest of the three
terms and reduces as the trailing edge Mach number drops,
becoming negative for high values of uncovered turning (Fig. 12).
The Mises trailing edge model is approximate but the results are
consistent with experimental observations, for example, Zhou
et al. [32] found a negative base pressure loss for the T106C blade
(a1 ¼ 32:7 deg and a2 ¼ 63:7 deg).
Fig. 14 Fully turbulent profile loss for the designs shown in
Fig. 2: (a) RANS; (b) Mises; (c) Eq. (7) for Mises
5 Endwall Loss
Figure 16 presents the endwall loss coefficients for the designs
in Fig. 2, which increase for the high-turning designs toward the
losses increase for designs with higher turning (bottom right) and bottom-right-hand corner of the plot. The level of endwall loss is
higher solidity (top right). lower than the profile loss (Fig. 14(a)) primarily due to the large
Considering a control volume downstream of a blade throat, aspect ratio of these designs (h=Cx ¼ 3). In order to understand
Denton [1] derived the following approximation for the mixed-out the trend in Fig. 16, the following analysis approximately sepa-
profile loss: rates the endwall loss into contributions from the two primary
!2
2ðhSSþPS Þ t þ dSSþPS Cpb t
nDenton  þ  (7)
s cos b2 s cos b2 s cos b2

where hSSþPS and dSSþPS , respectively, represent the sum of trail-


ing edge momentum and displacement thickness; b2 is the blade
metal exit angle; t is the trailing edge thickness; and Cpb is the
base pressure coefficient. The loss coefficient predicted by Eq. (7)
is presented in Fig. 14(c) for the Mises results. It shows a similar
trend to the calculated losses in Fig. 14(b), but there are differen-
ces in the overall levels of up to 18%. Michelassi et al. [33] per-
formed direct numerical simulations of unsteady flow around the
T106 blade and noted a discrepancy of around 10–15% between
the calculated “steady” loss sources and the terms in Eq. (7). The
current results demonstrate similar differences for steady Mises
simulations, indicating that the variance observed by Michelassi is
probably due to the assumptions underpinning Eq. (7), such as
zero deviation angle, rather than any unsteady phenomena. Fig. 16 Endwall entropy loss coefficients

081004-6 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


entropy-generating mechanisms: “background” dissipation in the calculate an isentropic edge Mach number using the inlet
endwall surface boundary layers and dissipation induced by the freestream conditions.
mixing of the secondary flows and their interaction with the sur-
The two methods give very similar trends, with the former
face boundary layers.
method estimating slightly higher losses. Compared to the overall
endwall loss in Fig. 16, it can be seen that both models predict a
similar level of endwall losses for the low-turning designs in the
5.1 Background Dissipation Loss. Considering the wetted- top-left of the plot. However, the overall trend is not captured, and
area of the endwalls, there is an inevitable entropy-generation rate the losses are underpredicted for higher-turning designs which are
simply due to the presence of boundary layers on these surfaces. likely to produce stronger secondary flows (e.g., see Ref. [34]).
This component of loss effectively represents a background dissi- For the remainder of this paper, the background dissipation loss
pation rate, i.e., the loss that would occur in the absence of the nCD in Eq. (10) is taken from the full integral method (Fig. 17(d))
blade secondary flows (as in the case of a bladeless duct). as this is believed to be more accurate. It should, however, be
Denton [1] described a simple model that assumed constant dis- noted that the simple method in Fig. 17(c) may be more useful
sipation coefficient in the boundary layer (CD  0:002). This from a preliminary design perspective.
gives an entropy-generation rate per unit area:

dS_ surf =dA ¼ CD qfs Vfs3 =Tfs (9)


5.2 Secondary-Flow-Induced Loss. Secondary flows induce
where dS_ surf =dA is the absolute entropy-generation per unit sur- entropy generation largely through the dissipation of secondary
face area, and Vfs is the velocity at the edge of the boundary layer. kinetic energy (ske) and the mixing-out of low streamwise-
For constant dissipation coefficient, this equation can be inte- momentum fluid. Figure 18(a) presents an estimate of the
grated to obtain a loss coefficient (after Eq. (1)): secondary-flow-induced losses, calculated by taking the difference
between the endwall loss nendwall (Fig. 16) and the simple model
 ð    3   of endwall boundary layer dissipation loss nCD described above
Aendwall T2 qfs Vfs A
nCD  2CD d (Fig. 17(d)):
h s cos a2 Tfs q2 V2 Aendwall
(10)
n secflow ¼ nendwall  nCD (11)
This component of loss is therefore inversely proportional to
aspect ratio. For the fixed aspect ratio of the current study, Fig. Figure 18(a) clearly shows the increase in losses for highly
17(a) presents the ratio Aendwall =ðh s cos a2 Þ, which relates the turning designs toward the bottom right of the plot. Squire and
area of the endwall to the flow cross-sectional area at exit; this Winter [34] produced a simple model of secondary flow in a cas-
factor is predominantly a function of the exit flow angle. The inte- cade of turning vanes by considering the convection of inlet
gral in Eq. (10) depends on the velocity ratio of the cascade boundary layer vorticity through an idealized blade passage with a
(Fig. 17(b)) and the velocities inside the blade passage. Two high axial-chord to pitch ratio. From their equations for the sec-
approaches have been taken to estimate this factor: ondary flow components (Eqs. (21) and (22) in Ref. [34]), it can
be shown that, for a given inlet profile, the normalized secondary
(1) The first method (Fig. 17(c)) follows Denton’s simple kinetic energy (ske) is proportional to
approach and uses information only from the blade pressure
distributions. The Mach numbers upstream and downstream "    #2
of the blade row are taken to be the freestream inlet and ske V1 s cos a2
/ ða1  a2 Þ (12)
exit values, respectively. Inside the passage, the freestream 0:5 V22 V2 Cx
Mach number is assumed to vary linearly in the tangential
direction between adjacent blade pressure and suction This factor is presented in Fig. 18(b). Although it captures the
surfaces. trend of the CFD results to some extent, it does not capture the
(2) In the second approach (Fig. 17(d)), the endwall static pres- peak in loss for the highest turning designs in the bottom-right-
sure is extracted from each RANS calculation and used to hand corner of Fig. 18(a).
Hawthorne [2] presented a more sophisticated vortex-filament
analysis to predict the strengths of the passage and counter vorti-
ces (Fig. 1) as well as the trailing edge shed vorticity. The method
predicts the orientation of the downstream vortex filament by con-
sidering the difference in transit times of flow passing over the
pressure and suction surfaces. Nondimensionally, this time differ-
ence may be written as

Fig. 17 Estimated background dissipation loss: (a) ratio of


endwall and exit flow area; (b) midspan velocity ratio; (c) simple Fig. 18 (a) Estimated contributions from secondary-flow-
estimate; (d) integrated estimate induced loss and (b) the SKE factor of Ref. [34] (Eq. (12))

Journal of Turbomachinery AUGUST 2017, Vol. 139 / 081004-7

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


ð     ð     2   3
   V2 S V2 S  V2 
DT ¼ TPS  TSS ¼ d  d (13)  2 6   sin a1  sin a2  7
PS Vfs Cx SS Vfs C x V1 4 DT Cx  V  5
AFPV ¼ M þ 1 (18)
V2 s cos a2 cos a2
where Vfs is the freestream velocity at the edge of the boundary
layer. The Hawthorne analysis predicts stronger secondary flows    !
when the pressure surface velocity is low and is therefore sensitive V1 
 V1 DT Cx
ðM  1Þ  V2 
AFCV ¼ M þ  sin a2  sin a2 
to the significant variation in pressure-side velocities around the V2 V2 s cos a2 
cos a2 V1 
design space (Figs. 3 and 10(b)). This sensitivity can also be
understood with reference to the basic flow physics: Since low (19)
velocity fluid on the pressure surface lacks streamwise momen-    
tum, it is more susceptible to the cross-passage pressure gradients, V1 M  V1 

AFSHED ¼  sin a  sin a
cos a2  V2 
2 1
and thus, the secondary flow will be stronger. V2
Hawthorne derived relationships for the streamwise circulation   
of the distributed passage vortex (PV), counter vortex (CV), and ðM  1Þ  V2 
  sin a2  sin a1  (20)
trailing edge shed vortex (SHED). These equations are here pre- 
cos a2 V1 
sented as nondimensional vorticity amplification factors (AFs),
which represent the ratio of outlet streamwise vorticity to the inlet where the compressibility factor M is given by
boundary layer vorticity:  
  c1 2
V2 h M ¼ 1 þ M1 (21)
C sec = s cos a2 2
Cx 2
AF ¼ (14) For the current designs, this factor varies between about 1 and
x1 =ðV1 =Cx Þ
1.04, indicating relatively modest compressibility effects in line
where C sec is the secondary circulation at the cascade outlet, with the findings of Marsh [35].
which is normalized using the exit velocity and flow area of the Figure 20 shows the magnitude of Marsh’s amplification factors
half-passage cascade. For the constant aspect ratio and boundary (Eqs. (18)–(20)) and their summation. As inspection of the above
layer height employed in this study, the inlet vorticity x1 is sim- equations shows that the distributions of the passage and counter
ply proportional to the inlet velocity V1 . vortex factors are almost identical to those of Hawthorne. However,
Hawthorne’s amplification factors may be expressed as (after the shed vorticity factor is markedly different, as discussed by Came
Eqs. (26), (32), and (37) in Ref. [2]) and Marsh [36]. The overall summation in Fig. 20(d) has the same
trend as the estimated secondary-flow-induced loss in Fig. 18(a).
0  1
 V2  Figures 21 and 22 present the quantitative comparisons of the
 2 B   sin a1  sin a2 C estimated secondary-flow-induced loss and the summed
V1 @ DT Cx  V1 A
AFPV ¼ þ (15)
V2 s cos a2 cos a2
 2
V1 DT  Cx
AFCV ¼  (16)
V2 s cos a2
 
V1 cos a1
AFSHED ¼  j tan a2  tan a1 j (17)
V2 cos a2

The magnitude of these three factors is plotted in Fig. 19, together


with the summation. Immediately, it can be seen that the passage
vortex factor and the summation approximately follow a similar
trend to the estimated secondary-flow-induced loss in Fig. 18(a), Fig. 20 Vorticity amplification factors, after Marsh [35]: (a) pas-
reinforcing the assertion that these losses are driven by the sec- sage vortex, (b) counter vortex, (c) shed vorticity, and (d) total
ondary flows.
Using a different approach, Marsh [35] applied Kelvin’s circu-
lation theory and included the effect of compressibility. The
amplification factors are calculated by integrating Marsh’s expres-
sions for vorticity (Eqs. (42), (51), and (63) in Ref. [35]) across
the exit flow area to obtain the total circulation:

Fig. 19 Vorticity amplification factors, after Hawthorne [2]: (a)


passage vortex, (b) counter vortex, (c) shed vorticity, and (d) Fig. 21 Comparison of estimated secondary-flow-induced loss
total and the summed amplification factor of Hawthorne [2]

081004-8 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


While there is a definite trend with the Hawthorne analysis in
Fig. 21, the near-proportional correlation with the Marsh amplifi-
cation factor in Fig. 22 is compelling. The secondary-flow-
induced-loss approximately collapses onto the trend:

nMARSHfit  0:0021 AFMarsh (22)

where the summed amplification factor AFMarsh is given by


2   3
 V2 
 2 6   sin a1  sin a2  7
V1 4 DT C x
V  5
1
AFMarsh ¼ 2M þ (23)
V2 s cos a2 cos a2

There is some scatter from this best fit line, of the order of
60.0015 for most designs but rising significantly for the designs
with the highest secondary-flow-induced losses (>0.008).
It should be noted that despite its satisfying simplicity and
accuracy, the fit in Eq. (22) is empirical and further work is
Fig. 22 Comparison of estimated secondary-flow-induced loss
and the summed amplification factor of Marsh [35] needed to understand its physical basis. The increase in the scatter
for designs with large secondary flows also needs to be under-
stood. Preliminary investigations suggest that some of this
amplification factors of Hawthorne and Marsh. The designs from scatter is an artifact of the finite aspect ratio used in the study
Fig. 2 are labeled (Tmax ¼ 0:2Cx ), and around 120 additional blade (h=Cx ¼ 3). Even at this relatively large aspect ratio, strong sec-
designs have been included in these plots. These include a set of ondary flows will induce some streamtube contraction at midspan
35 designs matching those in Fig. 2, but with lower maximum and hence affect the breakdown of profile and endwall loss in Eq.
thickness ðTmax ¼ 0:15Cx Þ and trailing edge wedge angle (6 deg). (2). Marsh’s theory may also be less applicable to these designs,
The surface Cp distributions for these thinner designs are shown particularly since it does not account for the significant stream
in Fig. 23 and are very similar to those presented in Fig. 3. The surface twist. Finally, it is noted that in this study, the inlet bound-
reduction in the passage blockage causes two main differences ary layer thickness has been scaled with the blade axial chord,
with the thicker set: there is a less pronounced overspeed on the which is likely to be somewhat arbitrary when considering the
early suction surface for the a1 ¼ 40 deg and a2 ¼ 50 deg impact of inlet conditions on the secondary flow. It will remain
design; and the pressure-side velocities are slightly lower, an for future investigations to examine these effects.
effect captured by the simple model in the Appendix. Also
included in Figs. 21 and 22 is a survey of different Cp design
styles with the original thickness distribution, conducted by vary- 5.3 Overall Endwall Loss: A Simple Method. From the
ing diffusion factor (0.16, 0.28, and 0.40) and peak suction loca- above analysis, cascade endwall loss can be estimated by the sum-
tion (PSL) (0.42, 0.52, and 0.62). These designs approximately mation of two components:
correspond to the parametric set of high-lift low-pressure turbine (1) Background Dissipation Loss nCD (Eq. (10)): based on a
designs studied by Coull et al. [18] and are labeled by the diffu- constant dissipation coefficient (CD ¼ 0:002) and an esti-
sion factor (DF) in Figs. 21 and 22. mate of the freestream velocity above the endwall boundary
layers (described in Sec. 5.1, Fig. 17(d)).
(2) Secondary-Flow-Induced Loss nMARSHfit (Eq. (22)):
which is approximately proportional to the vorticity ampli-
fication factor after Marsh [35] (Eq. (22); Fig. 22).
Figure 24 compares the result of this method with the calcu-
lated endwall losses for all of the designs studied (around 150 in
total). The average discrepancy is 5.0%, with most of the cases
falling within 610% of the correlated values.
The method can be generalized for cascades with different
aspect ratios: the endwall boundary layer dissipation in Eq. (10)
already has the correct sensitivity, while the secondary-flow-
induced loss should be scaled with the inverse of aspect ratio. The
method is only strictly valid for the inlet condition studied (collin-
ear turbulent boundary layer with h=Cx ¼ 0.01), and further work
is needed to extend the approach to different inlet conditions.

6 Conclusions
A parametric design study has been conducted to examine the
impact of key design variables on endwall loss in turbine cas-
cades. The impact of flow angles, blade thickness, and suction sur-
face aerodynamics has been studied and the following conclusions
are drawn:
 The pressure-side velocity is largely determined by the
cross-passage pressure gradient in the covered turning
region, being lower for higher-turning and thinner designs.
Fig. 23 Mises Cp distributions for the thinner designs: (a) rela- The variation affects both the blade lift coefficient and the
tive to outlet pressure and (b) relative to trailing edge pressure endwall losses.

Journal of Turbomachinery AUGUST 2017, Vol. 139 / 081004-9

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


Cp ¼ pressure coefficient ¼ ðP01  PÞ=ðP01  P2 Þ
Cx ¼ axial chord
Cpb ¼ base pressure coefficient ¼ ðPbase  P2 Þ=ðP01  P2 Þ
CpTE ¼ Cp referenced to TE ¼ ðP01  PÞ=ðP01  PTE Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
DF ¼ diffusion factor ¼ ðCpTE ðpeakÞ  1Þ
h ¼ span, or specific enthalpy (Eq. (1) only)
M ¼ Mach number
P, P0 ¼ static and total pressure
PSL ¼ peak suction location ¼ sSS ðpeakÞ=s0SS
Recx ¼ axial chord Reynolds number ¼ V2 Cx =t
ReS0 ¼ suction surface Reynolds number ¼ V2 S0SS =t
s ¼ pitch (spacing), or specific entropy (Eq. (1) only)
S ¼ distance along surface
S0 ¼ total surface length
t ¼ trailing edge thickness
T ¼ temperature
Tmax ¼ maximum blade thickness
T  ¼ nondimensional surface transit time
V ¼ velocity
x ¼ axial distance
Fig. 24 Endwall loss: simple model versus CFD Yp ¼ total pressure loss coefficient
Zw ¼ Zweifel lift coefficient
a ¼ flow angle
b ¼ metal angle
 The Zweifel lift coefficient is highly dependent on flow C ¼ circulation
angles and is not a universal measure of blade loading. For t ¼ kinematic viscosity
designs with matching suction surface Cp distributions, the n ¼ entropy loss coefficient (see Eq. (1))
suction surface contribution to Zweifel varies by up to 40% q ¼ density
as the flow angles change. A lift coefficient based on circula- x1 ¼ mass-averaged inlet boundary layer vorticity
tion [14] allows a more consistent comparison across the
design space.
 Endwall loss may be separated into the contributions of: (1) Subscripts and Abbreviations
dissipation in the endwall boundary layers and (2) losses
induced by the secondary flow, which includes the dissipa- CD ¼ endwall boundary layer dissipation
tion of secondary kinetic energy. fs ¼ boundary layer edge
 Secondary-flow-induced loss scales almost linearly with a LE ¼ leading edge
vorticity amplification factor derived using the method of mid ¼ midspan
Marsh [35]. mix ¼ mixed-out (constant-area)
 A simple model is presented which captures the calculated peak ¼ peak suction location on suction surface
endwall loss for designs with different flow angles, thickness PS ¼ pressure surface
distributions, and suction surface Cp distributions. The aver- SS ¼ suction surface
age discrepancy between modeled and calculated endwall TE ¼ trailing edge
loss is around 5%, with higher scatter for designs with stron- 0¼ stagnation
ger secondary flow. 1, 2 ¼ row inlet and outlet
 The model developed is based on physical principles with a
single correlation coefficient. The method achieves good
accuracy for the cases studied, which span a significant pro- Appendix: Pressure-Side Loading Model
portion of the design space, and thus provides a suitable basis In order to understand the trends observed in pressure-side
for a low-order prediction of cascade endwall loss. loading (Figs. 8(b) and 10(b)), Fig. 25 presents a simple model for
Further work is needed to examine the effects of inlet condi- the covered turning region (after Hodson [37]). The flow is
tions and aspect ratio and to relate cascade endwall loss to real assumed to rotate as a free vortex between two circular arcs which
turbine behavior. approximate the suction and pressure sides of adjacent blades.
From the geometry shown, the radii of curvature are
Acknowledgment
The author would like to acknowledge the funding, support,
and permission to publish from Rolls-Royce plc; in particular to
Frederic Goenaga, Raul Vazquez, and Giulio Zamboni. The
author also thanks Howard Hodson, John Denton, Graham Pullan,
Chris Clark, and Shaun Kaufmann of the Whittle Laboratory for
their input.

Nomenclature
A¼ area
AF ¼ vorticity amplification factor (Eq. (14))
AR ¼ aspect ratio ¼ h=Cx Fig. 25 Free-vortex model for the covered turning region: (a)
CD ¼ dissipation coefficient Þ pffiffiffiffiffiffi geometric approximation, (b) free-vortex, and (c) pressure
Co ¼ circulation coefficient ¼ Cp dðS=S0SS Þ, Ref. [14] distribution

081004-10 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


References
[1] Denton, J. D., 1993, “Loss Mechanisms in Turbomachines,” ASME J. Turbo-
mach., 115(4), pp. 621–656.
[2] Hawthorne, W. R., 1955, “Rotational Flow Through Cascades—Part 1: The
Components of Vorticity,” Q. J. Mech. Appl. Math., 8(3), pp. 266–279.
[3] Sieverding, C. H., 1985, “Recent Progress in the Understanding of Basic
Aspects of Secondary Flows in Turbine Blade Passages,” ASME J. Eng. Gas
Turbines Power, 107(2), pp. 248–257.
[4] Langston, L. S., 2001, “Secondary Flows in Axial Turbines—A Review,” Ann.
N. Y. Acad. Sci., 934(1), pp. 11–26.
[5] Adrian, R. J., Christensen, K. T., and Liu, Z.-C., 2000, “Analysis and Interpreta-
tion of Instantaneous Turbulent Velocity Fields,” Exp. Fluids, 29(3),
pp. 275–290.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi [6] Denton, J. D., and Pullan, G. P., 2012, “A Numerical Investigation Into the
Fig. 26 Free-vortex model of CP2PS for the thin (a) and thick Sources of Endwall Loss in Axial Flow Turbines,” ASME Paper No. GT2012-
(b) sets of blades 69173.
[7] Pullan, G. P., Denton, J. D., and Dunkley, M., 2003, “An Experimental and
Computational Study of the Formation of a Streamwise Shed Vortex in a Tur-
bine Stage,” ASME J. Turbomach., 125(2), pp. 291–297.
[8] Ainley, D. G., and Mathieson, G. C. R., 1957, “A Method of Performance
Estimation for Axial-Flow Turbines,” ARC Reports and Memoranda No.
2974.
[9] Dunham, J., and Came, P. M., 1970, “Improvements to the Ainley–Mathieson
Method of Turbine Performance Prediction,” ASME J. Eng. Power, 92(3),
pp. 252–256.
[10] Kacker, S. C., and Okapuu, U., 1982, “A Mean Line Performance Method
for Axial Flow Turbine Efficiency,” ASME J. Eng. Power, 104(1), pp.
111–119.
[11] Craig, H. R. M., and Cox, H. J. A., 1971, “Performance Estimation of Axial
Flow Turbines,” Proc. Inst. Mech. Eng., 185(1), pp. 407–424.
[12] Traupel, W., 1977, Thermische Turbomaschinen, Springer-Verlag, Berlin.
[13] Benner, M. W., Sjolander, S. A., and Moustapha, S. H., 2006, “An Empirical
Prediction Method for Secondary Losses in Turbines—Part II: A New Second-
Fig. 27 Pressure-side circulation integral for the thin (a) and ary Loss Correlation,” ASME J. Turbomach., 128(2), pp. 281–291.
[14] Coull, J. D., and Hodson, H. P., 2013, “Blade Loading and Its Application in
thick (b) sets of blades, from Mises design calculations
the Mean-Line Design of Low Pressure Turbines,” ASME J. Turbomach.,
135(2), p. 021032.
[15] Sharma, O. P., and Butler, T. L., 1986, “Predictions of Endwall Losses and Sec-
RPS ¼ Cx =ðsin a1  sin a2 Þ (A1) ondary Flows in Axial Flow Turbine Cascades,” ASME J. Turbomach., 109(2),
pp. 229–236.
[16] Drela, M., 1985, “Two-Dimensional Transonic Aerodynamic Design and Anal-
RSS ¼ Cx =ðsin a1  sin a2 Þ  ðs  Tmax Þ (A2) ysis Using the Euler Equations,” Ph.D. thesis, Massachusetts Institute of Tech-
nology, Cambridge, MA.
A limit of RSS  0 is applied to prevent nonphysical solutions in [17] Youngren, H., and Drela, M., 1991, “Viscous/Inviscid Method for Preliminary
some cases. For simplicity, the flow is taken as incompressible. Design of Transonic Cascades,” AIAA Paper No. 91-2364.
[18] Coull, J. D., Thomas, R. L., and Hodson, H. P., 2010, “Velocity Distributions
Considering the pressure gradient across the passage (Fig. 25(b)), for Low Pressure Turbines,” ASME J. Turbomach., 132(4), p. 041006.
one can obtain [19] Goodhand, M. N., and Miller, R. J., 2011, “Compressor Leading Edge
Spikes: A New Performance Criterion,” ASME J. Turbomach., 133(2), p.
dP qV 2 2  021006.
¼  P0  Pð yÞ (A3) [20] Howell, R. J., Ramesh, O. N., Hodson, H. P., Harvey, N. W., and Schulte, V.,
dy R Rð yÞ 2001, “High Lift and Aft-Loaded Pressure Profiles for Low-Pressure Turbines,”
ASME J. Turbomach., 123(2), pp. 181–188.
   2 [21] Vazquez, R., Torre, D., Partida, F., Arman~anzas, L., and Antoranz, A., 2011,
P0  Pð yÞ V ð yÞ 2 RPS “Influence of Surface Roughness on the Profile and End-Wall Losses in Low
) ¼ ¼ (A4) Pressure Turbines,” ASME Paper No. GT2011-46371.
P0  PPS VPS RSS þ y
[22] Shahpar, S., and Lapworth, L., 2003, “PADRAM: Parametric Design and Rapid
Meshing System for Turbomachinery Optimisation,” ASME Paper No.
Integrating across the passage, the mass flow per unit span can be GT2003-38698.
equated to the value at exit. Using the result in Eq. (A4) [23] Moinier, P., and Giles, M. B., 1998, “Preconditioned Euler and Navier–Stokes
Calculations on Unstructured Grids,” 6th ICFD Conference on Numerical
ð sTmax   Methods for Fluid Dynamics, Oxford, UK, Mar. 31–Apr. 3.
RPS [24] Menter, F. R., Langtry, R. B., Likki, S. R., Suzen, Y. B., Huang, P. G., and
qV2 s cos a2 ¼ qVPS dy (A5)
0 RSS þ y V€olker, S., 2006, “A Correlation-Based Transition Model Using Local
Variables—Part I: Model Formulation,” ASME J. Turbomach., 128(3),
    pp. 413–422.
VPS pffiffiffiffiffiffiffiffiffiffiffiffi s cos a2 RPS [25] Halstead, D. E., 1996, “Boundary Layer Development in Multi-Stage Low Pres-
) ¼ CpPS ¼ =ln (A6) sure Turbines,” Ph.D. thesis, Iowa State University, Ames, IA.
V2 RPS RSS
[26] Benner, M. W., Sjolander, S. A., and Moustapha, S. H., 2006, “An Empirical
pffiffiffiffiffiffiffiffiffiffiffiffi Prediction Method for Secondary Losses in Turbines—Part I: A New Loss
Predictions of CpPS using this equation are shown in Fig. 26 Breakdown Scheme and Penetration Depth Correlation,” ASME J. Turbomach.,
for a range of flow angles and two different thickness distribu- 128(2), pp. 273–280.
[27] Hodson, H. P., and Dominy, R. G., 1987, “The Off-Design Performance of
tions. Figure 26(a) shows predictions for the set with maximum a Low-Pressure Turbine Cascade,” ASME J. Turbomach., 109(2), pp.
thickness Tmax ¼ 0:15Cx , discussed in Sec. 5.2. Figure 26(b) 201–209.
shows the results for the thicker designs in Fig. 2. In general, the [28] Gregory-Smith, D. G., Graves, C. P., and Walsh, J. A., 1988, “Growth of Sec-
pressure surface velocity increases for designs with low deflection ondary Losses and Vorticity in an Axial Turbine Cascade,” ASME J. Turbo-
mach., 110(1), pp. 1–8.
and larger thickness. The predictions can be compared with the [29] Zweifel, O., 1945, “The Spacing of Turbomachine Blading, Especially With
pressure-side integral of the circulation coefficient (Eq. (6)) for Large Angular Deflection,” Brown Boveri Rev., 32(12), pp. 436–444.
the two sets of design, presented in Fig. 27. It can be seen that the [30] Coull, J. D., and Hodson, H. P., 2012, “Predicting the Profile Loss of High-Lift
overall trend with flow angles and thickness is captured, but the Low Pressure Turbines,” ASME J. Turbomach., 134(2), p. 021002.
[31] Curtis, E. M., Hodson, H. P., Banieghbal, M. R., Denton, J. D., Howell,
free-vortex model predicts a lower value since it does not capture R. J., and Harvey, N. W., 1997, “Development of Blade Profiles for
the rapid pressure surface acceleration toward the trailing edge Low-Pressure Turbine Applications,” ASME J. Turbomach., 119(3), pp.
(e.g., see Fig. 3). 531–538.

Journal of Turbomachinery AUGUST 2017, Vol. 139 / 081004-11

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab


[32] Zhou, C., Hodson, H. P., and Himmel, C., 2014, “The Effects of Trailing Edge [35] Marsh, H., 1976, “Secondary Flow in Cascades—The Effect of
Thickness on the Losses of Ultra-High Lift LP Turbine Blades,” ASME J. Tur- Compressibility,” Aeronautical Research Council R&M No. 3778.
bomach., 136(8), p. 081011. [36] Came, P. M., and Marsh, H., 1974, “Secondary Flow in Cascades: Two Simple
[33] Michelassi, V., Chen, L. W., Pichler, R., and Sandberg, R. D., 2015, Derivations for the Components of Vorticity,” J. Mech. Eng. Sci., 16(6),
“Compressible Direct Numerical Simulation of Low-Pressure Turbines—Part pp. 391–401.
II: Effect of Inflow Disturbances,” ASME J. Turbomach., 137(7), p. 071005. [37] Hodson, H. P., 2012, “Blade to Blade Flowfields in Axial Turbomachines,”
[34] Squire, H. B., and Winter, K. G., 1951, “The Secondary Flow in a Cascade of Cambridge Turbomachinery Course 2012, Vol. 1, University of Cambridge,
Airfoils in a Nonuniform Stream,” J. Aeronaut. Sci., 18(4), pp. 271–277. Cambridge, UK, pp. 35–75.

081004-12 / Vol. 139, AUGUST 2017 Transactions of the ASME

Downloaded From: http://turbomachinery.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jotuei/936123/ on 03/22/2017 Terms of Use: http://www.asme.org/ab

You might also like