You are on page 1of 8

46

Comparison of analytical and numerical predictions of stator


vane pressure distribution
produced by mean rotor wake impingement
Hélène Posson1*, Frédéric Sicot2, Stéphane Moreau1, Nicolas Gourdain3

1
Groupe d’Acoustique de l’Université de Sherbrooke (GAUS), Département de Génie Mécanique,
Université de Sherbrooke
2500 Boulevard de l’Université, Sherbrooke, QC, J1H4X2, CANADA

Phone: +1-819-821-8000 ext:62085, FAX: +1-819-821-7163, E-mail: helene.posson@usherbrooke.ca
2
ONERA DADS/MS, Châtillon, FRANCE
3
CERFACS, Computational Fluid Dynamics Team, Toulouse, FRANCE
A BST R A C T methods: an Unsteady Reynolds Average Navier Stokes Simulation
A recently developed analytical model for the predictions of the (URANS) and a Harmonic Balance Method (HB). The two simu-
fan tonal and broadband noises has been tested by comparing the lations are used to outline the ability of the HB to accurately predict
unsteady blade loading with results provided by numerical un- the detailed unsteady aerodynamics as it has already been shown to
steady simulations on a realistic compressor rotor-stator configu- predict mean aerodynamic performances quite well while reducing
ration (CME2). Both a blade slice and the actual full blade span the computational time.
have been investigated with the URANS and Harmonic Balance The analytical model and the numerical simulations are first
methods. In the simpler case of the blade slice, both numerical presented. Then a comparison is performed in a narrow annulus
methods compare very well up to the sixth harmonics, but show configuration both with the real shape of the vanes and with flat plat
large fluctuations and variations close to the leading edge caused by vanes. Finally, the comparison is performed in the fully annular
a small local flow separation, which cannot be predicted by the configuration to outline the three-dimensional effects and the abil-
analytical model and yield larger levels on the first and second ity of the model to capture these phenomena.
harmonics. In the full 3D case, the better flow field behaviour on
the blade triggers a better agreement between the URANS and the A N A LY T I C A L M O D E L A N D N U M E R I C A L SI M U L AT I O NS
analytical model. A further comparison with a flat plate cascade
improves the comparison for the first harmonics. A nalytical model
The noise prediction model is a strip-theory approach (Posson
INTRODUC T ION and Roger, 2008) based on a previously published formulation for
Modern turbofan engines have a larger high bypass ratio in the unsteady blade loading in rectilinear cascade (Posson and Roger,
order to improve the aircraft performance while diminishing the fan 2007). The latter follows Glegg's analytical work (1999) for the
rotational speed. Then the jet noise has been reduced and the fan response of a swept blade row to a three-dimensional gust. This
noise becomes another important contribution to the total noise compressible linear unsteady model provides a continuously valid
particularly at approach condition. In the latter conditions, the fan analytical formulation for the velocity potential produced by a
rotational speed is lower, and the blades and vanes operate under three-dimensional gust impinging on a swept blade row of infinite
subsonic speeds for which the unsteady loading on the blades and span. That is obtained resorting to the Wiener-Hopf technique and
vanes constitute the main noise sources. As pointed out by Prasad extensively using the residue theorem.
& Verdon (2002), the unsteady aerodynamic phenomena on blades
and inside the cascade also have a critical role in the outbreak of Rectilinear cascade model. The blade row modelled in Fig. 1
sustained force vibrations, flutter and consequent fatigue of fan, is taken as an equivalent rectilinear cascade of infinite span made of
compressor or gas turbine. Besides, the turbofans have a rather high infinitely thin flat plates of finite chord c. The blades can be stag-
solidity, compared to ventilators, which will affect both their gered at an angle  and swept at an angle .
aerodynamic and acoustic performances. Therefore accurate pre-
diction schemes, involving notably cascade and three-dimensional
geometry effects, become essential for the design of quieter fans. In
preliminary industrial designs, analytical models can still provide
faster results than numerical simulations of the real unsteady
compressible three-dimensional flow around the actual fan and
OGV geometries. These time-marching unsteady simulations, even
inviscid, are still computationally memory and time-consuming.
The present paper is the first part of a twofold study aimed at as-
sessing a recently developed analytical model for the prediction of
unsteady vane loading (Posson and Roger, 2007). It focuses on the
impingement of mean rotor wakes in the context of tonal noise Fig. 1 Geometry of the rectilinear cascade: (a) side view; (b)
predictions. This unsteady vane loading is then used as an equiva- three-dimensional view.
lent dipole source distribution in the framework of an acoustic
analogy. The study addresses a comparison with two numerical The mean flow U0 (vectorial sum of the chordwise Uc and
spanwise direction Wc component) is assumed at a zero an-
Copyright © 2010 by ISROMAC-13
-1-
gle-of-attack, i.e. parallel to the blade surface. The amplitude of the The three-dimensional cascade configuration is then simulated by
incident gust, representing a fluctuation around the mean flow is dividing the blade row in annular strips along the span. Each strip is
supposed to be small enough to justify the linearization of the flow unwrapped and assimilated to an equivalent rectilinear cascade of
field equations, a Fourier decomposition of the incident gust and a infinite spanwise extent with perfectly thin and rigid flat plates,
recombination of the cascade response by a superposition principle. having the local geometrical parameters of the blade row. The mean
This three-dimensional gust (of radial wave number kzc) can be flow and the incident wake excitation parameters are deduced in the
either an acoustic wave propagating downstream or upstream with reference frame of the rectilinear cascade from those of the three
an angle  or a vortical wave convected downstream by the mean dimensional case at the studied radius. Then the unsteady blade
flow. Solely, the latter is investigated here to deal with the rotor loading, produced by a three dimensional gust is predicted by the
wakes/stator interaction. analytical model. This technique enables to take into account radial
The unsteady blade loading formulation is obtained by variations of the blade-row geometry. The unsteady blade loading
re-addressing the expression for the velocity potential  for the distribution on the blades is finally considered as the equivalent
scattered field of the incident gust in the reference strip 0 yc h source of sound in the usual sense of the acoustic analogy. The
given by Eq. (1) (Glegg, 1999) wave equation for the sound field is solved using the Green's
function of the annular rigid duct with a uniform mean flow. The
1 

~ e i y c e  i y c 
 ( xc , y c , z c , t )   
2 1  e ih  ik xc d  i

1 e  ih  ik xc d  i 

... (1) three-dimensional cascade configuration is then simulated with a
reasonable accuracy, expecting faster computations than with fully
D ( k xc )e ik xc xc ii t  ik zc z x dk xc three-dimensional numerical simulations. The model has been first
applied to the fourth category of the second Computational
where the inter-blade phase angle , the incident angular velocity
AeroAcoustics (CAA) benchmark (Namba and Schulten, 2000).
i and kzc characterize the incident gust excitation,
The unsteady blade loading and the tonal noise produced by a
kxc=(i-kzcWc)/Uc is the chordwise wavenumber and D is the Fou- harmonic gust on an annular cascade are in a very good agreement
rier transform of the discontinuity of the velocity potential  across with those of fully three-dimensional annular models in a narrow
the n=0 blade and its wake. D can be decomposed as the sum of the annulus configuration (Posson and Roger, 2007, Posson et al, 2009).
four terms D(i) corresponding respectively to (1) the direct scatter- Yet, large discrepancies observed in a low hub-to-tip ratio case have
ing of the incident gust, (2) the trailing-edge back-scattering, and (3 led to the introduction of a correction of the blade loading analytical
and 4) the last two coupled scattering corrections by the leading and formulation (Posson and Roger, 2007) to account for the actual
trailing edges leading to the exact solution. Each D(i) is expressed wave equation in an annular duct.
analytically according to the Wiener-Hopf technique. After comparing the numerical results for a harmonic gust, the
Glegg provides the analytical formulation for the acoustic field next step is to compare the results of the analytical model in an
outside the blades by calculating (1) with the residue theorem, the actual stage configuration. The incident excitation is no longer a
acoustic field inside the blades being computed numerically by harmonic gust but an actual rotor wake. The latter is either meas-
Boquilion et al. (2003). Posson & Roger (2007) have recently ured (Envia, 2002) or computed from URANS (Gourdain and
extended it to the velocity potential field inside the cascade and Leboeuf, 2009) or HB simulations (Sicot, 2009).
then to the pressure distribution on the blades (by finding the left
and right limits) thanks to an extensive use of the residue theorem Unsteady Reynolds Average Navier Stokes Simulation
and a splitting of the inter-blade area in three areas in a staggered In the URANS simulations, the Reynolds averaging is an en-
cascade case (Fig.2). semble averaging of flow realizations. Yet to capture the low fre-
quency unsteady flow phenomena, the averaged quantities remain
time dependent yielding the Unsteady RANS model. The averaging
process removes the random turbulent phenomena of the flow,
which are modeled by moments of different order. Various turbu-
lence models coexist and provide closure to the RANS equations.

H ar monic Balance M ethod


While the URANS model could simulate virtually any kind of
unsteady flows, the Harmonic Balance (HB, Gopinath et al., 2007)
(a) 0dc/2
method only addresses time-periodic flows for which the funda-
mental frequency is a priori known. It is a numerical method de-
rived from the semi-discrete equations of the URANS equations,
which are cast into a set of coupled steady problems. These steady
problems correspond to a uniform sampling of the flow within the
time-period and are simultaneously solved. They are coupled by a
specific spectral high-order time-derivative operator only valid for
time-periodic flows. The steady problems are solved using classical
RANS techniques such as a pseudo-unsteady time marching to
(b) c/2dc reach the steady state. They benefit from convergence acceleration
techniques such as local time stepping, multigrid or implicit algo-
Fig. 2 Splitting of the inter-blades space into patches (a,b,c) and rithms. They all converge to a "snapshot" of the flow at their given
segments (I,II, III) or (I, II’, III) to express analytically the pressure time instant, and, as they are converged simultaneously, the HB
jump on the reference blade. method can be considered parallel in time.

The unsteady blade loading formulation takes into account a Simulations parameters
three-dimensional incident gust impinging on a three-dimensional In the present study, all CFD computations are performed with
rectilinear cascade and has been validated by comparisons with the ONERA solver elsA (Cambier and Veuillot, 2008) which is a
alternative models (Smith, 1973; Whitehead, 1987; Atassi and cell-centered multi-block structured solver. The spatial discretiza-
Hamad, 1981). tion is carried out by the 2nd order Jameson-Schmidt-Turkel scheme
This unsteady blade loading distribution is then meant to be (Jameson et al., 1981). A dual time-stepping method is used and a
used in a full 3D annulus configuration by a strip theory approach. convergence study leads to employ 160 time steps per time-period

-2-
and 20 sub-iterations. The latter use a backward Euler scheme the tip and hub ignored in the analytical model. Excellent agree-
where the implicit problem is solved by an LU-SSOR scheme ment was also found between the above two numerical approaches
(Yoon and Jameson, 1987). The single equation turbulent model of for both the mean and instantaneous flow field and the overall blade
Spalart and Allamaras (1992) is applied in conjunction with wall performances. For instance, Figs. 4 and 5 show the instantaneous
laws. pressure and axial velocity fields in the blade passage at 75% of the
A uniform flow is injected at inlet and a throttle condition with slice span (or 55% of the full blade span) for the HB simulation.
radial equilibrium is prescribed at outlet. A no-slip wall condition is The pressure field stresses that the stagnation point is close to the
applied on blades and on the casing and the hub in the 3D case. In stator leading edge (incidence close to 0°) as the rotor wakes im-
the radial slice case, the radial boundaries are modeled by a slip pacts on the stator blade. Yet, some strong non-homogeneity of the
wall condition. For the azimuthal boundaries, the phase lag periodic flow can already be seen in the axial velocity contours. Similar
conditions (Erdos et al., 1977) allow to solve for only one blade contours are obtained with the URANS simulation.
passage per row.

The HB computations use the same numerical parameters as


URANS except that the implicit problem is solved by a BJ-SSOR
method (Sicot et al., 2008) to take the HB time-derivative operator
into account. The phase periodic condition is performed by a
spectral interpolation. As each row captures the blade passing
frequency of the opposite row, rotor/stator interactions requires the
coupling of two HB computations, one per row with its own fun-
damental frequency. Different time samplings are tried and they are
much coarser than that of URANS, ranging from 5 to 13 instants
per period.
The multi-block structured mesh is composed of 51 points in
the azimuthal direction. There are 183 points around the rotor blade
and 171 around the stator blade. The blade slice has 5 points in the
radial direction while there are 57 points for the full 3D case. The
tip gap is discretized by 10 points. A blade-to-blade cut in the 3D
mesh is shown in Fig. 3. Fig. 4 HB static pressure contours at 75% slice span (55% blade
span).

Fig. 3 Two-dimensional snapshot of the mesh around a rotor blade


and a stator vane.

B E N C H M A R K C O N F I G U R AT I O N Fig. 5 HB axial velocity contours at 75% slice span (55% blade


The benchmark configuration is the subsonic compressor span).
CME2 (Michon et al., 2005) with a rotor of 30 blades and a stator of
40 vanes with a hub-to-tip ratio H =0.716. In each studied confi- A more quantitative assessment of the unsteady loading on the
guration, the actual rotor wake, computed numerically by the HB or stator blade is obtained by performing a Fourier transform of the
the URANS solver is used as the incident excitation of the stator. It instantaneous wall pressure field. Fig. 6 compares the corre-
is then Fourier decomposed into a series of harmonic gusts which sponding amplitudes of the first three harmonics for the two
impinge on the stator vanes. The latter are used as incident gusts in simulations on the full slice. The amplitudes of the fluctuations are
the analytical model. The resulting unsteady vane loading is then decreasing and more variations are observed, with increasing fre-
calculated at the blade passing frequency and its first harmonics. quency as expected. Moreover for all frequencies, larger ampli-
tudes are observed at the stator leading edge where the wake is
N A R R O W A N N U L US R ESU LTS impacting. For all three cases, very good agreement is achieved
between the two numerical simulations. Yet, some significant radial
Real stator vanes: effect of vanes camber, thickness, and load- variations are observed for all frequencies. It was then found that,
ing unexpectedly, some flow separation (stressed by the larger ampli-
First, a computation is performed on a slice of the real com- tudes with two peaks on the blade pressure side of the fluctuations
pressor geometry described above. This preliminary stage is first harmonics in Fig. 6) was obtained at the hub despite the narrow
studied as this computation on a slice can be assimilated to a narrow annulus and the negligible variation of stagger angle in the slice.
annulus configuration where the analytical model was shown to Such a small radial non-equilibrium then triggers the radial varia-
yield good results. Moreover, Sicot (2009) used it as his main tions and the larger amplitude variations seen at the slice bottom.
turbomachinery-validation test-case of the HB method integrated in Such a flow phenomenon is not expected to be captured by the flat
the Onera’s solver ElsA. Slip boundary conditions selected at the plate analytical model.
top and bottom of the slice were also limiting the wall-effects near

-3-
levels and variations from 20 to 80% of the chord length. Similarly,
for the second harmonics (2 BPF), additional large fluctuating
humps are found around mid-chord. These large fluctuations could
be traced to the larger fluctuations close to the hub, moving radially
as shown in Fig 6. The more oscillating behaviour of the simula-
tions close to the leading edge in all harmonics is also a hint at the
effect of the flow separation at the bottom of the slice. The larger
oscillations at mid-chord might also be traced to the effect of
camber and thickness of the CME2 blade profile and the resulting
higher loading compared to a flat plate.
F lat plate vanes
To partially explain the discrepancies seen in the slice case, and
provide a better validation case for the analytical model, a slice of
flat plates without any variation of flow angle is also considered.
The resulting cascade has 30 vanes and the stagger angle is as-
sumed to be 16°. A fictitious wake is imposed as a time-evolving
(a) BPF (top: URANS; bottom: HB) boundary condition for the URANS simulation in order to impose a
mean flow that impinges the cascade at almost a zero an-
gle-of-attack (assumption of the analytical model). The outer-wake
mean flow has actually no incidence with respect to the flat plate
cascade. Whereas the angle-of-attack of the inner-wake mean flow
slightly departs from zero. Fig. 7 shows the shape of this wake
colored by the contours of axial Mach number. The resulting mul-
ti-block structured mesh of the flat plate cascade is plotted in Fig. 9.
It is composed of 51 points in the azimuthal direction. There are
191 points around the stator blade. The blade slice has 31 points in
the radial direction.

(b) 2BPF (top: URANS; bottom: HB)

Fig. 7 Fictitious wake model in the flat plate cascade. Contour


levels are the axial Mach number.

(c) 3BPF (top: URANS; bottom: HB)

Fig. 6 Wall pressure contours on the slice.

The comparisons with the analytical model are achieved on the


normalized pressure jumps at various radii for several harmonics.
Fig. 9 shows the real (figures (a) to (d)) and imaginary (figures (e)
to (h)) parts of the first four harmonics of the normalized pressure
jump as a function of the position along the axial chord length. The Fig. 8 Mesh of the flat plate cascade.
main contribution is found at the leading edge for all cases and the
overall levels are similar and would yield overall similar acoustic Again the normalized pressure jumps at various radii of the BPF
levels. This is all the more the case for the higher harmonics (third and the first two harmonics are compared between the URANS
and fourth here) for which the mean loading might have a slighter simulation and the analytical model in Fig. 10. Figures (a) to (c)
impact. Both the levels and the overall shapes are rather well cap- correspond to the real parts, and figures (d) to (f) to the imaginary
tured by the analytical model. For the BPF however, both the real parts.
and imaginary parts of the unsteady simulations show much higher

-4-
(a) BPF: real part (e) BPF: imaginary part

(b) 2 BPF: real part (f) 2 BPF: imaginary part

(c) 3 BPF: real part (g) 3 BPF: imaginary part

(d) 4 BPF: real part (h) 4 BPF: imaginary part


Fig. 9 Pressure jumps at 75% slice span (55% blade span).
Solid lines: URANS; triangles: HB; circles: Analytical model

-5-
At the BPF both results compare favourably for both real and
imaginary parts, even though oscillations are observed in the
simulation. The comparison gets worse with increasing frequencies
with higher and larger and larger oscillations appear at the trailing
edge, where a large peak is exhibited.
However the flat blades are still too thick and a significant
vortex shedding appears in the wake of the URANS simulations.
This in turn induces an additional source of pressure fluctuations
mainly concentrated at the trailing. The vortex shedding frequency,
based on a Strouhal number St = f e/U xd =0.2, where e is the plate
thickness, is equal to f =9750 Hz, which is near to the 3 BPF fre-
quency: 9495 Hz. Yet the vortex shedding modifies also smaller
frequencies, and above all the higher harmonics.

(d) BPF: imaginary part

(a) BPF: real part

(e) 2 BPF: imaginary part

(b) 2 BPF: real part

(f) 3 BPF: imaginary part


Fig. 10 Pressure jumps at 50% flat plate cascade.
Solid lines: URANS; circles: Analytical model

F U L LY 3D A N N U L US C O N F I G U R AT I O NS
The final comparison is achieved on the full stage configura-
tion of the CME2 set-up.
Fig. 10 again shows the real (figures (a) to (d)) and imaginary
(c) 3 BPF: real part (figures (e) to (h)) parts of the first four harmonics of the normal-
ized pressure jump as a function of the position along the axial
chord length. The URANS results are compared with the analytical
model. The model predicts the same overall levels as the simulation
for all four harmonics. Beyond the fourth frequency, the numerical
simulations do not have enough resolution and the levels drop
drastically. The differences are attributed, firstly, to the definition
of the incident gusts from the mean rotor wakes, and secondly, to

-6-
three dimensional effects. In order to conclude on the limitations of
the model concerning this last point, the authors first need to im-
prove the modelling of the wakes to feed correctly the model.

(e) BPF: imaginary part

(a) BPF: real part

(f) 2 BPF: imaginary part

(b) 2 BPF: real part

(g) 3 BPF: imaginary part

(c) 3 BPF: real part

(h) 4 BPF: imaginary part


Fig. 11 Pressure jumps at 55% blade span.
(d) 4 BPF: real part Solid lines: URANS; circles: Analytical model

-7-
Runge-Kutta Time-Stepping Schemes”, In AIAA 14th F luid and
C O N C L USI O NS Plasma Dyna mic Conference, Palo Alto, California , number AIAA
Both numerical unsteady simulations and analytical predictions Paper 81-1259.
have been achieved on a realistic rotor-stator configuration. Three Michon, G. J., Milton, H. and Ouayahya, N., 2005,  “Experi-
configurations have been considered: a blade slice, a flat plate and mental study of the unsteady flows and turbulence structure in an
the actual full blade span. On the case of the blade slice, the axial compressor from design to rotating stall conditions”, In 6th
URANS and HB methods compare very well for all harmonics European Turbomachinery Conference, Lille, F rance.
studied (up to the sixth harmonics), but show large fluctuations and Namba, M. and Schulten, J. B. H. M., 2000, “Category 4 – Fan
variations close to the leading edge, especially at the bottom of the stator with harmonic excitation by rotor wake”, In J. C. Hardin, D.
slice. These fluctuations and oscillations can be traced to a small Huff, and C.K. Ta m, edictors, Third Computational Aeroacoustics
local flow separation, which cannot be predicted by the analytical (CAA) Workshop on Benchmark Problems, NAS A Conference
model. Consequently, the levels are larger in the simulations than in Publication, CR- 2000-209790, pp. 73–86.
the analytical model with larger oscillations mainly on the first and Prasad, N. and Verdon, J. M., 2002, “A three-dimensional lin-
second harmonics. Yet, the overall differences in levels might not earized Euler analysis of classical wake/stator interactions: valida-
have a significant impact on the overall noise level of the stator tion and unsteady response predictions”, J. Aeroacoustics, Vol. 1
noise due to turbulent wake impact. On the case of the flat plate (2), pp. 137–163.
configuration, the analytical model provides a consistent result for Posson, H., and Roger, M., 2007,  “Parametric study of gust
the BPF and its first harmonics, yet a vortex shedding is observed scattering and sound transmission through a  blade  row”, 13th
by the URANS simulation leading to a large peak at the trailing AIAA/C EAS Aeroacoustics Conference and Exhibit, Rome, Italy,
edge and large oscillations for high harmonics. Finally, on the number AIAA Paper 2007-3690, pp. 1–22.
actual full blade configuration the overall behaviour is captured by Posson, H., and Roger, M., 2008, “Experimental validation of a
the analytical model. Yet further investigations are needed in par- cascade response function for fan broadband noise predictions”,
ticular on a proper definition of the rotor wakes seen as a superpo- 14th AIAA/C EAS Aeroacoustics Conference and Exhibit, Vancou-
sition of incident harmonic gusts in the analytical model. ver, Canada , number AIAA Paper 2008-2844, pp. 1–20.
Posson, H., Moreau, S. and Roger, M., 2009, “Classical
AC KNOWLEDGEM ENT wake/stator interactions addressed with an analytical cascade
We wish to acknowledge the support of SAFRAN-SNECMA model”, Canadian Aeronautics and Space Institute AERO’09 
for providing us the CME2 rotor-stator geometry, and the RQCHP Conference, Aerodyna mics Symposium, Ottawa, Canada, number
(Réseau Québécois de Calcul Haute Performance) Canada for 303, pp. 1–9.
providing the necessary computational resources. Sicot, F., 2009, “Simulation efficace des écoulements instation-
naires périodiques en turbomachines”, Thèse de Doctorat de
R E F E R E N C ES l’Ecole Centrale de Lyon, spécialité Mécanique, F rance, number
2009-19.
Atassi, H., and Hamad, G., 1981, “Sound generated in a cascade Sicot, F., Puigt, G. and Montagnac, M., 2008, “Block-Jacobi
by the three-dimensional disturbances convected in subsonic flow”, Implicit Algorithms for the Time Spectral Method”, AIAA J., Vol.
7th AIAA Aeroacoustics Conference, Palo Alto, California , number 46(12), pp.3080–3089.
AIAA Paper 1981-2046, pp. 1–13. Smith, S. N., 1973,  “Discrete frequency sound generation in
Boquilion, O., Glegg, S. A. L., Devenport, W. J. and Larsan, J., axial flow turbomachines,” Brit. Aeronaut. Res. Coun. R&M , 3709,
2003, “The interaction of large scale turbulence with a cascade of pp.1-59.
flat plates”, 9th AIAA/C EAS Aeroacoustics Conference and Exhibit, Spalart, P. R. and Allmaras, S. R., 1992, “A One-Equation
Hilton Head, South Carolina, number AIAA Paper 2003-3289, pp. Turbulence Transport Model for Aerodynamic Flows”, In 30th
1–8. AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada ,
Cambier, L. And Veuillot, J., 2008, “Status of the elsA Software number AIAA Paper 92-0439.
for Flow Simulation and Multi-Disciplinary Applications”, 46th Whitehead, D. S., 1987, “Classical two-dimensional methods”,
AIAA Aerospace Sciences Meeting and Exhibit, Reno,Nevada, AGARD Manual on aerolasticity in axial flow turbomachines,
number AIAA Paper 2008-0664. Vol.1, Unsteady Turbomachinery Aerodynamics, Chap. 3
Denton, J. D. and Singh, U. K., 1979, “Time Marching Methods (AGARD-AG-298), pp. 1-30.
for Turbomachinery Flow Calculation”,  In E. Schmidt, editor: Yoon, S. and Jameson, A., 1987, “An LU-SSOR Scheme for the
Application of Numerical Methods to F low Calculations in Tur- Euler and Navier-Stokes Equations”, In AIAA 25th Aerospace
bomachines, VKI Lecture Series. Von Kármán Institute for F luid Sciences Meeting and Exhibit, Reno Nevada , number AIAA Paper
Dyna mics, Rhode-St-Genèse (Belgium). 87-0600.
Envia, E., 2002, “Fan noise source diagnostic test – Vane un-
steady  pressure  results.”, NAS A Technical Memorandum,
TM-2002-211808.
Erdos, J. I., Alznert, E. and McNally, W., 1977, “Numerical
Solution of Periodic Transonic Flow through a Fan Stage”. AIAA J.,
Vol. 15(11), pp. 1559–1568.
Glegg, S. A. L., 1999, “The response of a swept blade row to a
three-dimensional gust”, J. Sound Vib., Vol 227 (1), pp. 29–64.
Gopinath, A., Van der Weide, E., Alonso, J. J., Jameson, A.,
Ekici, K. and Hall, K. C., 2007, “Three-Dimensional Unsteady
Multi-Stage Turbomachinery Simulations using the Harmonic
Balance Technique”, In 45th AIAA Aerospace Sciences Meeting
and Exhibit, Reno, Nevada , number AIAA Paper 2007-0892.
Gourdain, N. and Leboeuf, F., 2009, “Unsteady simulation of an
axial compressor stage with casing and blade passive treatments”,
AS ME J. Turbomach., Vol 131 (2) 021013 (12pages).
Jameson, A., Schmidt, W. and Turkel, E.,1981,  “Numerical
Solutions of the Euler Equations by Finite Volume Methods Using

-8-

You might also like