You are on page 1of 10

Proceedings of ASME Turbo Expo 2014: Turbine Technical Conference and Exposition

GT2014
June 16 – 20, 2014, Düsseldorf, Germany

GT2014-26813

AERODYNAMIC AND AEROACOUSTIC OPTIMIZATION OF A TRANSONIC


CENTRIFUGAL COMPRESSOR

Peng Wang Mehrdad Zangeneh ∗


Department of Mechanical Engineering Department of Mechanical Engineering
University College London University College London
Gower Street, London, WC1E 6BT Gower Street, London, WC1E 6BT
United Kingdom United Kingdom
Email: ucahpwa@ucl.ac.uk Email: m.zangeneh@ucl.ac.uk

ABSTRACT NOMENCLATURE
The performance of transonic compressors can be charac-
terized aerodynamically and aeroacoustically. In this paper, Roman symbols
the DLR SRV2 compressor without vaned diffusers and its re- c speed of sound
designed version are studied. The redesign strategy (Zangeneh f integral surface
et al. 2011 [1]) utilized the 3D inverse design and CFD analysis. fn shaft rotational frequency
Both compressors were analyzed in ANSYS CFX 11, and the FFT fast Fourier transform
computational results show that the predicted pressure-ratio and H( f ) Heaviside function
efficiency of the original compressor have good agreement with g = τ − t + r/c
experimental results. The simulations have also revealed that the G Green’s function
redesigned one is superior at both design and off-design points at M0 convective Mach number
different rotating speeds. n normal vector
This work applies a convective FW-H method to further in- p0 acoustic pressure
vestigate the noise radiation from these two compressors. As Pi j compressible stress tensor
the blade tip speed is supersonic, the permeable integral surface p0T thickness noise
scheme must be adopted. The flow quantities needed as the in- p0L loading noise
puts to the FW-H solver were extracted from the CFD solutions. r source-observer distance, |xx − y|
The numerical predictions of the noise SPLs at blade passing S p permeable surface
frequency and its harmonics match the experimental measure- t observer time
ments reasonably well. It is found that the original compressor Ti j Lighthill stress tensor
has significant variations of SPLs as the operating mass flow rate u fluid velocity
changes whereas the redesigned one has much slighter variations. v solid surface velocity
At peak efficiency the redesigned compressor has a lower noise x observer position
level. y source position
This study provides insights for the optimal design of a tran-
sonic compressor when good aerodynamic and aeroacoustic per- Greekq symbols
formance are both required. β 1 − M02
δ ( f ) Dirac delta function
∗ Address all correspondence to this author.

1 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


δi j Kronecker delta, δi j = 1 for i = j, otherwise δi j = 0 important to consider aeroacoustic noise emissions as part of the
π ' 3.1415926 . . . design process of high pressure centrifugal compressors.
σi j viscous stress tensor In this work, the DLR SRV2 centrifugal compressor is used.
τ source time The impeller’s main parameters are given in the Table 1. The
detailed flow field in this impeller has been investigated by us-
Subscript ing detailed L2F measurements at DLR, see Krain and Hoffman
0 ambient quantity (1998) [2] and Eisenlohr et al. (2002) [3]. Also its aeroacoustic
i, j vector components performance has been measured by Raitor and Neise (2008) [4].
n projection in the normal direction This impeller was subject to a detailed redesign study by di-
r projection in the radiation direction rect design method, see Krain and Hoffman (2007, 2008) [5, 6],
who made changes to meridional shape, rotor axial length, tip
Superscript diameter, exit wrap angles and blade angle distribution, in or-
0 acoustic perturbation quantity der to optimize the impeller geometry. The performance of the
˙ source-time derivative new impeller, renamed SRV4, was found to be 2 points better
than the baseline at design speed but showed lower performance
Other symbols at other part-load conditions. In a separate study Zangeneh et
 wave operator al. (2011) [1] applied a 3D inverse design method to redesign
¯
 convective wave operator the impeller geometry. In this redesign effort, the meridional
shape and normal thickness distribution of the original SRV2 im-
Acronyms peller was maintained and a new blade shape was designed by
APU auxiliary power unit specifying the blade loading distribution. The resulting impeller
BPF blade passing frequency was found to have better performance compared to the original
CFD computational fluid dynamics SRV2 impeller at all operating conditions. In the present paper,
DLR Deutsches Zentrum für Luft- und Raumfahrt (German the detailed CFD predictions for the two impellers are used to-
Aerospace Center) gether with a recently developed far field acoustic model based
FWH Ffowcs-Williams and Hawkings on Ffowcs-Williams and Hawkings (1969) approach [7] to in-
L2F laser-two-focus vestigate the effect of aerodynamic changes in the inverse design
LE leading edge impeller on its noise emissions.
LHS left hand side A comparison of the main impeller and splitter blade geom-
PS pressure surface etry is shown in Figs 1 and 2, respectively. One can notice that
PR pressure ratio a larger curvature appears at the LE of both redesigned main im-
RHS right hand side peller and splitter blades. The redesigned blades are also more
TCN tip clearance noise leaned at the TE.
SPL sound pressure level The full-bladed views are presented in Fig. 3. In Ref. [1],
SS suction surface CFD analysis confirmed the superiority of aerodynamic perfor-
TE trailing edge mances of the redesigned compressor.
The objective presented herein is twofold. First, the flow
field details will be further examined. Second, the acoustic noise
will be numerically predicted at the design speed.
1 INTRODUCTION
High pressure transonic centrifugal compressors are used in
helicopter engines, APUs and marine Diesel turbochargers. In 2 AEROACOUSTIC MODEL
many of these applications there is an increasing demand to im- The study of aerodynamically generated noise was pio-
prove the compressor efficiency at higher pressure ratios. This neered by Lighthill (1952) [8] and then extended by Curle (1955)
has resulted in an increasing trend towards higher tip speeds [9]. However, Lighthill’s work was to investigate the noise
and hence high Mach number flow in the inducer, which sub- due to turbulence and Curle’s was to study the acoustic field
sequently gives rise to strong local shocks. Not only would these due to stationary solid bodies. In the end of 1960s, Ffowcs-
shocks affect the compressor stage performance but they are also Williams and Hawkings [7] generalized their predecessors’ work
the main sources of buzz-saw noise observed at the intake. In and derived the well-known FW-H equation, which characterizes
many applications environmental legislation is placed on noise that the acoustic field constitutes three sources, namely, thick-
emissions, for example as in the case of International Maritime ness/monopole, loading/dipole, and quadrupole noise in a static
Organization limit on turbocharger noise. It is therefore rather medium. This theory was further generalized by Najafi-Yazdi

2 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


TABLE 1. SRV2 compressor parameters
Shaft speed 50000 rpm
Flow rate at peak efficiency 2.73 kg/s
for vaneless diffuser stage
Impeller tip radius 112 mm
Tip speed 586 m/s
Blade count full/splitter 13/13
LE hub radius 30 mm
FIGURE 2. Solid models of the original(LHS) and redesigned(RHS)
TE tip radius 78 mm splitter

Blade angle LE tip 26.5 deg


Blade angle TE 52 deg
Exit blade height 8.7 mm
Diffuser inclination against 13 deg
radial
Tip clearance 0.5 mm at inlet to 0.3 mm at
exit

FIGURE 3. Solid models of the whole original(LHS) and re-


designed(RHS) compressor

sure, and

2 2 2 2 i
¯ 2 = 1 ∂ − ∂ + 2 U0 j ∂ + U0iU0 j ∂
h
 , (2)
c2 ∂t 2 ∂ x2j c2 ∂t∂ x j c2 ∂ xi ∂ x j

FIGURE 1. Solid models of the original(LHS) and redesigned(RHS)


main impeller blade is the convective wave operator, c is the speed of sound, U 0 is
the convective flow velocity, and

et al. (2010) [10] by including the convective flow, and their Qi = ρ(ui +U0i − vi ) + ρ0 (vi −U0i ),
extended formulation of FW-H equation was employed in the Li j = ρui (u j +U0 j − v j ) + (p − p0 )δi j − σi j ,
present work. The convective FW-H equation is shown below.
Ti j = ρui u j − σi j + [(p − p0 ) − c2 (ρ − ρ0 )]δi j , (3)

 
¯ 2 [H( f )p0 ] = ∂ ∂ ∂
 +U0 j [Qi ni δ ( f )] − [Li j n j δ ( f )] where p0 is the ambient pressure, u and v are the fluid and solid
∂t ∂xj ∂ xi surface velocity, respectively, σi j is the viscous stress tensor, δi j
∂2 is Kronecker delta tensor and Ti j is the well-known Lighthill
+ [H( f )Ti j ], (1)
∂ xi ∂ x j stress tensor. If there is no convective flow, i.e., U 0 = 0, then
Eq. (2) reduces to the usual FW-H equation. This equation can
be solved in a similar matter as Farassat’s formulations. A brief
in which f is integral surface function, δ ( f ) is the Dirac delta derivation will be given below and more details can be found in
function, H( f ) is the Heaviside function, p0 is the acoustic pres- Ref. [10]

3 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


∂ tZ Z
δ (g) 3
2.1 Integral formulations − [Li j n j δ ( f )] d ydτ
The classical FW-H equation has the key feature that it is ∂ xi −∞ V 4πr∗
valid for the external flow embedded in unbounded space; thus, ∂2
Z t Z
δ (g) 3
it can be solved using free-space Green function, + [H( f )Ti j ] d y dτ, (11)
∂ xi ∂ x j −∞ V 4πr∗

δ (g) in which M0 is the convective flow Mach number. Moving the


G(xx,t; y , τ) = , (4)
4πr observer-time derivative inside the integral and converting the
spatial derivative to the temporal derivative. (see details in Faras-
in which sat [13] or Najafi-Yazdi et al. [10])

r
g = τ −t + , (5) Z h Q̇ n + Q ṅ i
i i i i
c 4π p0T (x,t) = ∗ 2
dS p
f =0 r (1 − MR ) τ
Z h ∂ r∗ Qi ni i
and x is the observer position, y is the source position, r is the rel- + − dS p
ative distance between source and observer, t is signal reception f =0 ∂ τ r∗2 (1 − MR )2 τ
time and τ is the signal emission time. However, a little modifica- h∂M Qi ni
Z i
R
+ dS p
tion must be introduced here to take account of the presence of a f =0 ∂ τ r∗ (1 − MR )3 τ
subsonic mean flow. It is assumed without losing generality that Z h r̃ Q̇ n + r̃ Q ṅ + r̃˙ Q n i
3 i i 3 i i 3 i i
the mean flow is along the positive x3 -direction. A 3-dimensional −M0 dS p
f =0 r∗ (1 − MR )2 τ
free-space Green’s function for the convective equation given by h ∂ r∗ r̃3 Qi ni
Z i
Blokhintzev (1956) [11] is +M0 dS p
∗2 2
f =0 ∂ τ r (1 − MR ) τ
Z h∂M r̃3 Qi ni i
δ (τ − t + r/c) R
G(xx,t; y , τ) = , (6) −M0 dS p
4πr∗ f =0 ∂ τ r∗ (1 − MR )3 τ
Z h r̃∗ Q n i
3 i i
−U0 ∗2
dS p , (12)
where f =0 r (1 − MR ) τ

−M0 (x3 − y3 ) + r∗
r= , (7)
β2 h r̃ L̇ n + r̃ L ṅ + r̃˙ L n i
1
Z
i ij j i ij j i ij j
4π p0L (x,t) = ∗ 2
dS p
cf =0 r (1 − MR ) τ
1
Z h ∂ r∗ Li j n j r̃i i
− dS p
c f =0 ∂ τ r∗2 (1 − MR )2 τ
q
r∗ = β 2 [(x1 − y1 )2 + (x2 − y2 )2 ] + (x3 − y3 )2 , (8)
1 h∂M Li j n j r̃i i
Z
R
+ dS p
c f =0 ∂ τ r∗ (1 − MR )3 τ
1
Z h L n r̃∗ i
ij j i
+ dS p , (13)
U0 c f =0 r∗2 (1 − MR ) τ
M0 = 1 − , (9)
c
where [ ]τ indicates the integrand is evaluated at the retarded time,
MR = vi r̂i∗ /c, and
q
β= 1 − M02 . (10)
x1 − y1 x1 − y1 x3 − y3
r̃1 = , r̃2 = β 2 , r̃3 = β 2 . (14)
r∗ r∗ r∗
The solution to the convective FW-H equation (Eq. (1)) is then
found as, The dS p in Eqs. (12) and (13 ) is the elementary area on the inte-
gral data surface, thus the quadrupole source term is legitimately
t
 Z
δ (g) 3 neglected. The algorithm of implementing this formulation will
Z
0 ∂ ∂
p (xx,t) = +U0 Qi ni δ ( f ) d y dτ
∂t ∂ x3 −∞ V 4πr∗ be discussed in the next subsection.

4 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


2.2 Algorithm for numerical implementation
The most common method of numerically evaluating the
retarded-time integrals is to use the mid-panel quadrature method
[14, 15]. Suppose Eq. (12) and (13) can be recast in the form,

I(yy, τ)
Z h i
4π p0 (x,t) = dS pi , (15)
f =0 r∗α1 (1 − MR )α2 τ

where I(yy, τ) is the re-organized source strength function de- FIGURE 4. Schematics of the nosie testing rig
pending on source position and time, and α1 , α2 are integers
which depend on the type of source, such as, monopole, dipole.
1 8 0
It can be approximated as,

h I (y ,t − r∗ /c) i
N 1 6 0
i i
4π p0 (x,t) ≈ ∑ ∗α1
i
δ S pi , (16)
r
i=1 i (1 − M Ri ) 2 τ
α
1 4 0

S P L (d B )
The surface S p is divided into N panels and the integrand Ii is
evaluated at the centre of every panel, yi at the retarded times 1 2 0

that are calculated by the equation τ = t − r/c. Although this


retarded-time dominant algorithm is robust and efficient, there is 1 0 0
an important problem hindering the applications for supersonic
conditions. In the presence of supersonic motions, one source
8 0
point can emit signals at different times and those signals arrive 0 1 0 0 0 0 2 0 0 0 0 3 0 0 0 0 4 0 0 0 0 5 0 0 0 0
at the observer simultaneously. This means that the retarded-time F re q u e n c y (H z )
equation can have more than one root, which is not very conve-
nient for such as the Newton-Ralphson root-finder algorithm. FIGURE 5. Numerical prediction of noise spectrum at 50000 rpm
An alternative method has been proposed by Casalino
(2003) [16] to replace the retarded-time algorithm so that the
multiple-roots problem can be avoided. Instead of selecting the of 157 mm. A microphone ring with four microphones placed 90
observer times in advance, the source time series can be pre- degrees to each other is mounted on the inner wall of the duct,
scribed first and then determine when the observer would receive and this ring is placed at a distance of 165 mm from the leading
the signals. The advanced/reception time is directly calculated edge of the impeller. The incoming flow is assumed to have a
using Eq. (17), and every single τ only corresponds to single t. uniform velocity normal to the cross-section of the duct. A more
However, a sequence of equally spaced source times would not detailed explanation of the experimental set-up is given in Raitor
yield a sequence of equally spaced observer times. As Eq. (17) and Neise’s paper [4].
suggests
3.2 Validation
r(τ) The acoustic pressure signals received at the four measure-
t =τ+ (17)
c ment points were computed in time domain and then converted to
the frequency domain using FFT algorithm. Figure 5 shows the
t and τ do not hold a linear relationship due to the term, r(τ)/c. circumferentially averaged predicted spectrum at experimental
The observer time history needs to be interpolated to provide the conditions. The compressor operated at the incoming flow rate
contribution at desired observer times in order for further sum- 2.55 kg/s and design speed 50000 rpm, producing a tip Mach
mations. number about 1.7. The measurement spectrum is shown in Fig.
6. A noticeable and significant feature is that the sound level
has very high magnitudes at the BPF (blade passing frequency,
3 NOISE SPECTRUM 10833 Hz) and its harmonics, which are also clearly higher than
3.1 Test facility their neighbouring components. In addition to these BPFs tones,
In Fig. 4 is shown the schematic view of the noise testing rig. there are also tonal components spreading over a range of har-
The compressor is directly connected to a duct with a diameter monics of the rotational frequency, which are known as the buzz-

5 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


RANS simulations using multiple blade passages with perturbed
(not cyclically symmetric) blade geometries are required to ob-
tain buzz-saw tones. In this study, we only modelled one blade
passage using steady-state CFD solver and assumed the perfect
circumferential periodicity of flow quantities in all other 12 pas-
sages. Thus, the strength variations of the shocks and their non-
linear propagations were not included, resulting in failing to cap-
ture the buzz-saw tones at the multiples of shaft frequency.
Another important feature that deserves particular attention
is that the measured spectrum has a “local maximum” before the
BPF, which cannot be seen in the predicted spectrum. If one
takes a closer scrutiny on the measured spectrum, one may find
that this local maximum occurs at the 5 fn ( fn is shaft rotational
frequency), which is the fifth engine order and very close to the
cut-on frequency of the duct. This phenomenon is likely due
FIGURE 6. Measured noise spectrum (see Fig. 6a in Ref. [4]) to the resonance between the fifth engine order and cut-on fre-
quency of the duct. For this study, the duct wall is not included
in the numerical implementation for the simplification purpose,
0 .8 5
and this explains why this “local maximum” phenomenon did not
0 .8 0 appear on the predicted spectrum. The purpose of using a free-
η

0 .7 5 in v e r s e space space aeroacoustic model is to ascertain if design trends


o r ig in a l are favourably reducing noise emissions as intended rather than
0 .7 0
get very accurate predictions of noise field.
2 .5 2 .6 2 .7 2 .8 2 .9 3 .0
M a s s flo w ( k g /s ) The sound generated by a compressor that operates at dif-
1 5 2
ferent inlet flow conditions is distinct. Therefore, it is necessary
to compare the radiated noise levels at the entire operating range
1 5 0
so as to get a comprehensive understanding of aeroacoustic be-
S P L (d B )

in v e r s e
1 4 8
o r ig in a l haviours. Figure 7 compares the aerodynamic and aeroacoustic
1 4 6 performances of the original and redesigned compressor work-
1 4 4
ing from surging to choking conditions under full speed. The
2 .5 2 .6 2 .7 2 .8 2 .9 3 .0
SPLs are the tonal magnitudes at the BPF. It is interesting to note
M a s s flo w ( k g /s )
that the sound level of the redesigned compressor has slight vari-
FIGURE 7. Aerodynamic and aeroacoustic performance comparisons ations at the entire operating ranges, while the original compres-
between the original and redesigned compressor sor produces sound with significant changes as the mass flow
varies. More importantly, it can be seen that sound levels gener-
ated by the redesigned are lower than those by the original when
saw noise. The dominant source of these buzz-saw components the working mass flow rate is in range from 2.72 to 2.88 kg/s. In
is the rotor-alone pressure of shocks attached to the supersoni- fact, the computational work presented in Ref [1] confirmed that
cally rotating blade tips, and these shocks can propagate against the peak efficiency with the vaneless stage occurs around 2.73
the incoming flow to the upstream in the inlet duct. kg/s. Thus, the redesign compressor shows lower noise emission
By comparing Figs. 5 and 6 it can be seen that the tones at around its peak efficiency point. However, the baseline impeller
BPF and its harmonics are captured in prediction without large has lower level of noise at lower mass flow rates and close to the
discrepancies, but there are no buzz-saw components on the pre- stall point.
dicted spectrum. This is because the buzz-saw tones are due to In the following section, the CFD predictions of the flow
the amplitudes variations and nonlinear propagations of shocks physics in the blade channel at close to the peak efficiency point
near the blade tips, which would induce the energy redistribu- will be presented so as to explain the reason why the inversely
tion of the frequency spectrum, i.e., the initial energy at BPF redesigned compressor has a better aeroacoustic performance.
and its harmonics is redistributed at the harmonics of rotor shaft
frequency (engine orders). A more detailed explanation of buzz-
saw noise generation mechanism can be found in the work by 4 CFD analysis
McAlpine, Fisher and Tester (2006) [17]. In a similar fan noise The flow in the original SRV2 impeller was analyzed by us-
prediction study, Gliebe et al. (2000) [18] reported that steady ing commercial CFD CFX11 by Zangeneh et al. [1]. For this

6 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


0 .9
Is e n tr o p ic c o m p r e s s io n
e ffic ie n c y ( to ta l- to ta l)

0 .8
in v e r s e
0 .7
o r ig in a l

0 .6
2 .4 2 .5 2 .6 2 .7 2 .8 2 .9 3 .0
M a s s flo w ( k g /s )

6
( to ta l- to ta l)

5
in v e r s e
o r ig in a l
4
P R

3
2 .4 2 .5 2 .6 2 .7 2 .8 2 .9 3 .0
M a s s flo w ( k g /s )
FIGURE 10. Mach number contour comparison between the original
FIGURE 8. Comparison of the characteristics of the original and re- (LHS) and redesign (RHS) at 25% of span
design compressor

FIGURE 11. Mach number contour comparison between the original


FIGURE 9. Mach number contour comparison between the original (LHS) and redesign (RHS) at 50% of span
(LHS) and redesign (RHS) at 15% of span

have particularly chosen an operating point of 2.72 kg/s mass


flow for comparisons. The details of flow physics are given in
purpose a mesh independence study was carried out and a mesh
Figs. 9-13, which are the blade-to-blade views of the Mach num-
size of 880k nodes was eventually used for each blade channel.
ber contours at different spanwise positions ranging from 15%
The CFD prediction was then compared with the detailed L2F
(near the hub) to 95% (near the shroud). In Figs. 14-18 are
measurement of the flow inside the impeller and good correlation
shown the comparisons of pressure distributions in the stream-
was obtained between the predictions and measurements. Also
wise direction on the main blades and splitters. By observing
they found good correlation between predicted and measured
these figures, a few important insights can be obtained and they
pressure ratio for SRV2 impeller with vaneless diffuser tested
explain why the inversely designed compressor is superior.
at 100%, 80% and 60% speeds. The inverse designed impeller
was found to have higher pressure ratio and isentropic efficiency 1. Flow separations at the LE are clearly present on the origi-
at all speeds. nal compressor from the lowest spanwise sections whereas
Figure 8 revels a comparison of characteristics at the design they are absent on the redesign until 75% of the span, and
speed. In order to show how the improved flow field in the blade as Figs. 12 and 13 depict, the flow separation in main blade
passages is associated with this aerodynamic performance, we and splitter channel of inverse design is much weaker than

7 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


3 0 0
in v e r s e
2 0 0 o r ig in a l

P re s s u re (k P a )
1 0 0

-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n
3 0 0
in v e r s e
2 0 0
o r ig in a l

P re s s u re (k P a )
1 0 0

-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n

FIGURE 12. Mach number contour comparison between the original FIGURE 14. Streamwise blade loading distributions on the main
(LHS) and redesign (RHS) at 75% of span blade (lower) and splitter (upper) at 15% of span

3 0 0
in v e r s e
2 0 0 o r ig in a l

P re s s u re (k P a )
1 0 0

-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n
3 0 0
in v e r s e
2 0 0 o r ig in a l
P re s s u re (k P a )

1 0 0

-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n

FIGURE 13. Mach number contour comparison between the original FIGURE 15. Streamwise blade loading distributions on the main
(LHS) and redesign (RHS) at 95% of span blade (lower) and splitter (upper) at 25% of span

that in the original design. The reduction of flow separation from 0.2 to 0.4 streamwise position.
increases the effective passage area, and resultantly weaken-
ing the shock strengths. Figure 19 shows that the hub-to-shroud Mach number con-
2. Shocks at the LE of the main blade of the redesigned com- tour details at the leading edges of two compressors. It confirms
pressor are significantly weaker than those of the original up that shocks near the hub of the redesign are weaker than the orig-
to 50% of the span (Figs. 9 - 11), and it can also be found inal one; however, this phenomenon did not continue all the way
in Figs. 14-16 that the original main blade sustains stronger to the shroud. The shocks caused by the redesign after 50% of
loading near the LE up to the mid of span. the span have larger strength than those of the original. This is
3. Shocks formed between the PS of main blade and SS of further validated by Fig. 20, which shows the efficiency varia-
splitter in the redesigned compressor are weaker those in the tions at the TE from hub to shroud. It can be clearly seen that
original one at all spanwise positions. This is also confirmed the inverse design has higher efficiencies up to 60% of the span.
by the blade loading distributions. It can be seen in Figs. 9 Figure 21 demonstrates the efficiency and static entropy varia-
- 13 that the pressure loading near the LE of the inverse de- tions of two compressors from the impeller inlet to outlet. It can
signed splitter is smaller than that of the original splitter, and be seen that the redesigned impeller generated less static entropy
this phenomenon is quite obviously shown in Figs. 16 to 18 than the original design, which consequently results in a higher

8 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


3 0 0
in v e r s e 3 0 0
2 0 0 o r ig in a l in v e r s e
P re s s u re (k P a )

2 0 0 o r ig in a l

P re s s u re (k P a )
1 0 0
1 0 0
0
0
-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
-1 0 0
S tr e a m w is e lo c a tio n 0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
3 0 0
S tr e a m w is e lo c a tio n
in v e r s e 3 0 0
2 0 0 o r ig in a l in v e r s e
P re s s u re (k P a )

2 0 0 o r ig in a l

P re s s u re (k P a )
1 0 0
1 0 0
0
0
-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
-1 0 0
S tr e a m w is e lo c a tio n 0 .0 0 .2 0 .4 0 .6 0 .8 1 .0

S tr e a m w is e lo c a tio n
FIGURE 16. Streamwise blade loading distributions on the main
blade (lower) and splitter (upper) at 50% of span FIGURE 18. Streamwise blade loading distributions on the main
blade (lower) and splitter (upper) at 95% of span
3 0 0
in v e r s e
2 0 0 o r ig in a l
P re s s u re (k P a )

1 0 0

-1 0 0
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n
3 0 0
in v e r s e
o r ig in a l
2 0 0
P re s s u re (k P a )

1 0 0

-1 0 0
FIGURE 19. Mach number contour comparison between the original
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
(LHS) and redesign (RHS) at LE
S tr e a m w is e lo c a tio n

FIGURE 17. Streamwise blade loading distributions on the main


1 .0
blade (lower) and splitter (upper) at 75% of span

0 .8
efficiency.
The CFD simulations show that the redesigned impeller gen-
S p a n n o r m a lis e d

0 .6
erates weaker shocks in the inducer and significantly reduces the in v e r s e
flow separation. The weaker shock structure noticeably reduces o r ig in a l
the generated noise. 0 .4

0 .2
5 Conclusion
In this paper, we examined the aerodynamic and aeroacous-
tic performance of a transonic centrifugal compressor and its re- 0 .0
0 .6 0 .7 0 .8 0 .9 1 .0
designed version. The employed 3D inverse design methodology Is e n tr o p ic c o m p r e s s io n e ffic ie n c y
has provided a better solution in terms of the aerodynamic and
partially aeroacoustic aspects. The aerodynamic performance is FIGURE 20. Comparison of spanwise variation of isentropic com-
achieved for the entire operating range. Noise radiations can also pression efficiency at the TE
be reduced when the redesigned compressor was operating near

9 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1 .0 0
[7] Ffowcs Williams, J.E. and Hawkings, D.L., 1969. Sound
Is e n tr o p ic c o m p r e s s io n

0 .9 5 generated by turbulence and surfaces in arbitrary motion.


Philos Trans R Soc A, 264, pp.321-342.
e ffic ie n c y

0 .9 0
[8] Lighthill, M.J. 1952. On sound generated aerodynami-
in v e r s e
0 .8 5
o r ig in a l cally. I: general theory. Proc. R. Soc. London. A, 231,
0 .8 0 pp.564-587.
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n [9] Curle, N. 1955. The influence of solid boundaries upon
4 0
aerodynamic sound. Proc. R. Soc. London. A, 211, pp.505-
S ta tic e n tr o p y ( J k g ^ - 1 K ^ - 1 )

in v e r s e
2 0 o r ig in a l 514.
0
[10] Najafi-Yazdi, A., Bres, G.A., and Mongeau, L. 2010. An
acoustic analogy formulation for moving sources in uni-
-2 0
formly moving media. Philos Trans R Soc A, 467, pp.144-
-4 0 165.
0 .0 0 .2 0 .4 0 .6 0 .8 1 .0
S tr e a m w is e lo c a tio n
[11] Blokhinstev, D. 1956. Acoustics of a nonhomogeneous
moving medium. NACA TM-1399.
FIGURE 21. Comparison of isentropic compression efficiency in [12] Farassat, F. 1975. Theory of noise generation from moving
streamwise location from impeller inlet to outlet bodies with an application to helicopter rotors. tech. rep.,
NASA TR R-451, 1975.
[13] Farassat, F. 2007. Derivation of Formulations 1 and 1A of
peak efficiency. In the subsequent work, an optimization strategy Farassat. NASA TM-2007-214853, 2007
that couples inverse design method and the current aeroacoustic [14] Brentner, K.S. and Farassat, F. 2003. Modelling aerody-
model will be employed to redesign another transonic compres- namically generated sound of helicopter rotors. Progress
sor with vaned diffusers. in Aerospace Sciences, 36, pp.592-594.
[15] Wang, P. and Zangeneh, M. 2013. Numerical prediction of
aerodynamic sound generated in a supersonic centrifugal
ACKNOWLEDGMENT compressor. 19th AIAA conference, May 27 - 29, 2013,
We would thank Napier Turbochargers Ltd and EPSRC for Berlin, Germany.
their financial support of this project. [16] Casalino, D. 2003. An advanced time approach for acous-
tic analogy predictions. Journal of Sound and Vibration,
261, pp. 583-612.
REFERENCES [17] McAlpine, A., Fisher, M.J., and Tester, B.J. 2006. “Buzz-
[1] Zangeneh, M., Amarel, N., Daneshkhah, K., and Krain, saw” noise: A comparison of measurement with predic-
H. 2011 Optimization of 6.2:1 Pressure Ratio centrifugal tion. Journal of Sound Vibration, 290, pp.1202-1233.
compressor impeller. Proceedings of ASME Turbo Expo [18] Gliebe, P., Mani, R., Shin S., Mitechell, B., Ashford, G.,
2011: Power for Land, Sea and Air, June 6-11, 2011, Van- Salamah, S. and Connell, S. 2000. Aeroacoustic predic-
couver, Canada, GT-2011-46505. tion codes. NASA/CR-2000-210244.
[2] Krain, H. and Hoffman, B. 1998. Flow physics in high
pressure ratio centrifugal compressors. ASME-Summer
Meetings, FEDSM98-4853. 1998
[3] Eisenlohr, G., Krain, H., Richter, F. and Tiede, V. 2002. In-
vestigations of the flow through a high pressure ratio cen-
trifugal impeller. ASME Turbo Expo 2002, Amsterdam,
Netherlands, GT-2002-30394.
[4] Raitor, T. and Neise, W. 2008. Sound generation in cen-
trifugal compressor. Journal of Sound and Vibration, 314,
pp.491-509.
[5] Krain, H. and Hoffman, B. 2007. Improved High Pressure
Ratio Centrifugal Compressor. ASME Turbo Expo 2007,
Montreal, Canada, GT2007-27100.
[6] Krain, H. and Hoffman, B. 2008. Flow study of a
redesigned high-pressure-ratio centrifugal compressor.
Journal of Propulsion and Power, 24, pp.1117-1123.

10 Copyright © 2014 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like