Flutter Analysis
Flutter Analysis
IGTI 2014
June 16-20, 2014, Düsseldorf, Germany
GT2014-25474
NOMENCLATURE
a Speed of sound.
ABSTRACT A Area vector.
The objective of this paper is to describe an accurate and c Blade chord length.
efficient reduced order modeling method for aeroelastic (AE) Cp Specific heat at constant pressure.
analysis and for determining the flutter boundary. Without los- d structural deformation vector.
ing accuracy, we develop a reduced order model based on the E Specific total energy.
Volterra series to achieve significant savings in computational Fm Modal force vector.
cost. The aerodynamic force is provided by a high-fidelity so- F Inviscid flux vector.
lution from the Reynolds-averaged Navier-Stokes (RANS) equa- Fv viscous flux vector.
tions; the structural mode shapes are determined from the finite hm (n) mth -order Volterra series kernel.
element analysis. The fluid-structure coupling is then modeled K Stiffness matrix.
by the state-space formulation with the structural displacement L Length measured in the x-direction.
as input and the aerodynamic force as output, which in turn acts M Mass matrix.
as an external force to the aeroelastic displacement equation for M∞ Freestream Mach number.
providing the structural deformation. NASA’s rotor 67 blade is p Pressure.
used to study its aeroelastic characteristics under the designated Pκ (ω ) Turbulence production in the κ (ω ) equation.
operating condition. First, the CFD results are validated against Pr Prandtl number.
measured data available for the steady state condition. Then, qi Heat flux vector.
the accuracy of the developed reduced order model is compared
q∞ Dynamic pressure(=ρ∞ a2∞ )
with the full-order solutions. Finally the aeroelastic solutions
U Conservative variables.
of the blade are computed and a flutter boundary is identified,
r Position vector.
suggesting that the rotor, with the material property chosen for
Re Reynolds number.
the study, is structurally stable at the operating condition, free of
S Source term.
encountering flutter.
T Temperature.
U∞ Inflow velocity magnitude.
∗ Currently
ui Cartesian velocity components.
School of Mechanical and Aerospace Engineering, Queen‘s Uni-
versity, Belfast, UK.
x, y, z Axial (horizontal), spanwise and vertical direction.
This
c material is declared a work of the U.S. Government and is not sub- δi j Kronecker delta.
ject to copyright protection in the United States. Approved for public release; ξ,η Modal displacement.
distribution is unlimited.
1 Copyright
c 2014 by ASME
κ Turbulence kinetic energy. and a structural dynamics code, as a result they in turn determine
μ Viscosity. the reliability and efficiency of the tool.
ω Specific turbulence dissipation rate. For aerodynamics analysis, developments in computational
Ω Rotating speed. fluid dynamics over the past several decades have provided in-
σ Value of step function. creasingly powerful and reliable capabilities. The complex-
τ Pseudo time. ity and fidelity, hence its range of applicability, of analysis is
τi, j Viscous stress tensor. strongly correlated with the fast evolution of computer power:
V Control volume. from the early linearized potential flow solution to the current
Subscripts large eddy simulations using Navier-Stokes equations. Linear
A Aerodynamic. models are still used widely in the design phase. But develop-
∞ Far upstream or “free” stream. ments in computer technology and CFD methods and software
m Modal coordinate have made use of high-fidelity models feasible even in early stage
s Structural dynamics. of a design cycle. However, large eddy simulations are still far
t Total (stagnation) condition. too costly and from being timely to be adopted in the design pro-
turb Turbulence. cess.
In the current study, we employ the Reynolds-average
Navier-Stokes equations for which the turbulence terms are
1 INTRODUCTION closed with the two-equation κ -ω model, specifically the shear
NASA is considering new generations of aircraft that meet stress transport (SST) version by Menter [2]. The second-order
aggressive economic, noise and environmental targets; a spe- backward differencing is used for time-discretization. The non-
cial configuration, called N3-X, employs all electric power and linear inviscid terms are approximated by the AUSM+ -up [3]
propulsion systems by which the thrust force for the vehicle is method while the viscous terms approximated by the usual cen-
generated exclusively with an array of fans housed in compart- tered formulas. The resulting implicit algebraic system is then
mentalized flow paths. Hence, the designing of fans to meet es- solved by the LUSGS method [4].
sential considerations is paramount. Specifically, increasing per- For structural dynamics, one may invoke the full finite ele-
formance and operating life and reducing weight to optimize the ment analysis, as employed in the aeroelastic study of rotor 67
economic objective, while reducing noise and emission to meet by Doi [5]. The resulting fluid-structure system is a time depen-
environmental regulations. In pursuit of higher performance of a dent set of equations describing not only the flow variables in the
compressor/fan, the past design trend is to run at higher pressure entire domain, but also the motion of the structure immersed in
ratios and higher mass flow rates, thus moving close to flutter the fluid. The system can be solved either in the frequency [6] or
boundaries associated with surge or choke as defined in the com- time domain [5]. The frequency domain approach may be pre-
pressor map. Hence, it is important to ensure the compressor ferred for linear problems for its computational efficiency; how-
is structurally sound over the entire operating range, from the ever for a nonlinear problem, it is more efficient and accurate to
choke to the stall conditions. Structural vibration, either caused arrive at solution with the time domain approach.
by natural resonance or forced response, is a major consideration The time-domain computation for flutter analysis can be-
in assessing the devise’s structural integrity. The fluid-induced come costly when a large number of time-dependent solutions
instability of a compressor blade is typically not of concern un- of the fluid-structure system are needed, for example as part of
less it is tuned to the natural vibration frequency. However, it a design process. It is therefore desirable to reduce the com-
becomes an issue in transonic speed regime, because a small dis- putational cost by a significant factor, for example by at least
turbance can result in a large amplitude variation and nonlinear an order of magnitude or more. This can be readily achieved
behavior. The unsteady excursion of a shock wave through the by employing strategies called model order reduction (MOR), of
blade-to-blade passage can intermittently choke or stall the flow, which the harmonic balance, proper orthogonal decomposition,
potentially crossing the flutter boundary. The unsteady forces and Volterra series are among the most popular. [7] Model order
resulting from the shock motion are shown to have either stabi- reduction should not only save computational effort, but also re-
lizing or destabilizing effects, depending on the shock structure tain the fidelity of the original (full) system. This goal has been
and inter-blade phase angle. [1] well realized for linear problems through model reduction, but
With advances in computers and computational fluid dynam- not yet universally for nonlinear problems.
ics (CFD), aeroelastic analysis is fast becoming common for real For nonlinear problems, the Volterra series expansion is used
world designs. To be useful and adopted in practice, a compu- to approximate the input-output relationship of a nonlinear time
tational tool must be reliable for predicting the aeroelastic char- dependent system, with a capability of capturing ”memory” ef-
acteristics and just as importantly be efficient (cheap and fast). fects. This input-output concept is well suited for the aeroelastic
This computational tool will consist of an aerodynamics code analysis in which the aerodynamic force and structural deforma-
2 Copyright
c 2014 by ASME
tion can be formulated in this framework. Furthermore, flowfield and turbulence generation.
and structural dynamics have different time scales and their in-
teractions often respond with delay in time, i.e., with ”memory” ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
ρ ρ (ui − ugi ) 0
effects. The Volterra series has been applied in various fields ⎜ ρ u1 ⎟ ⎜ ρ (ui − ug )u1 ⎟ ⎜ pδi1 ⎟
of engineering and is mostly used to construct a reduced order ⎜ ⎟ ⎜ i ⎟ ⎜ ⎟
⎜ ρ u2 ⎟ ⎜ ρ (ui − ug )u2 ⎟ ⎜ pδi2 ⎟
model to mimic a complex dynamic system. Unsteady aerody- ⎜ ⎟ ⎜ i ⎟ ⎜ ⎟
U=⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ ρ u3 ⎟ , Fi = ⎜ ρ (ui − ugi )u3 ⎟ , Pi = ⎜ pδi3 ⎟ ,
namic force responses to wing motion have been calculated by ⎜ ρE ⎟ ⎜ ρ (ui − ug )E + pui ⎟ ⎜ 0 ⎟
Silva [8] using the Volterra theory. In the present work, we de- ⎜ ⎟ ⎜ i ⎟ ⎜ ⎟
⎝ ρκ ⎠ ⎝ ρ (ui − ugi )κ ⎠ ⎝ 0 ⎠
scribe the application of the Volterra series, based on RANS so-
ρω ρ (ui − ugi )ω 0
lutions, to turbomachinery aeroelastic problems. (2)
The paper is organized as follows. Section 2 gives the and the terms attributed to viscous diffusion
fluid and structure equations employed and outlines the methods
⎛ ⎞ ⎛ ⎞
adopted to solve them, especially including detailed description 0 0
of fluid-structure coupling and model order reduction based on ⎜ τi1 ⎟ ⎜ 0 ⎟
⎜ ⎟ ⎜ ⎟
the Volterra series. Section 3 presents the application to aeroelas- ⎜ τi2 ⎟ ⎜ −ρ u3 ⎟
⎜ ⎟ ⎜ ⎟
tic analysis for NASA’s rotor 67 compressor blade along with the Fvi = ⎜
⎜ τi3
⎟ , S = Ω ⎜ ρ u2 ⎟ ,
⎟ ⎜ ⎟ (3)
validation of the CFD solution against measured data. In Section ⎜ τi j u j + qi ⎟ ⎜ 0 ⎟
⎜ ⎟ ⎜ ⎟
3.2 we show the results of applying the developed reduced-order ⎝ qκi ⎠ ⎝ Pκ ⎠
model to find the flutter boundary of rotor 67. qωi Pω
∂uj
τi j = (μ + μturb ) ∂ ui
∂ x j + ∂ xx − 23 ∂∂ uxk δi j
μ μturb ∂ T
k (5)
qi = −( Pr + Prturb ) ∂ xi
d
UdV + (F + P) · dA = Fv · dA + SdV (1)
dt V ∂V ∂V V
The turbulence eddy viscosity for the study presented herein
is provided through the solution of two transport equations of
scalar quantities (κ , ω ), specifically the so-called κ -ω turbu-
where we have the standard notation for the conservative vari-
lence model [10] is enhanced with Menter’s shear-stress trans-
ables plus the turbulence variables in U. The surface integral on
port (SST) model [2]. The details of the turbulence model, well-
∂ V (x,t) consists of fluxes through the vectorial area dA can be
known and elaborated in the cited reference, are omitted here.
expressed in terms of 3 Cartesian coordinates. The relative con-
vective flux Fi , the pressure flux Pi , and the viscous stresses and
heat flux Fvi in the i-direction, i = 1, 2, 3 are given in Eq. (2-3), 2.2 CFD Solution Methods
written in the relative coordinate system moving with the speed To eliminate accumulative time integration error, we opt for
ug . [9] The source terms includes the rigid-body rotation, Eq. (4) the dual-time stepping approach, in which a time rate of change
3 Copyright
c 2014 by ASME
of U in pseudo time (τ ) is added to Eq. (1) so that at each determination of the modal (generalized) force Fm = {Fmi ; i =
new time level the unsteady CFD equations are balanced, namely 1, · · · , Nm }, where each component of the modal force is the in-
driving its discretized residual to diminish. The dual time step- ner product of the mode shape Φi and aerodynamic force vectors
ping strategy is formulated as: over the entire CFD nodes. Additionally, the physical deforma-
tion d of a structure can be expressed in terms of the mode shape
and the modal displacement ξ from the above dynamics equa-
tion, leading to physical displacement, d = ∑Ni=1m
Φi ξi and physi-
d
UdV = R(U) Nm
dτ V cal force f = ∑i=1 Φi Fmi .
The above second-order differential equation can be recast
d
=− UdV − (F + P) · dA + Fv · dA + SdV.
(6) into the following first-order differential system:
dt V ∂V ∂V V
In the physical time step (t) the residual R(U) is discretized η̇ = As η + Bs Fm (8)
implicitly using a 3-level, backward differencing in order to ob-
tain second-order temporal accuracy, resulting in a highly nonlin-
where
ear system. In the pseudo-time step the implicit system is solved
by performing fixed-point iterations untill the residual R(U) of
the nonlinear physical-time equation has diminished or reduced ξ 0 I 0
η= , As = , Bs = (9)
to specified small values. Then the solution is advanced to the ξ̇ −M−1 K 0 −M−1
next time level. This pseudo-time iteration is carried out by em-
ploying the LUSGS method [4]. and η = {ηi ; i = 1, · · · , Nm }.
The inviscid flux terms F + P arguably have received the The time derivative in Eq. (8) is approximated by the
most attention in past CFD research, especially those under the second-order Crank-Nicolson method, producing a discrete sys-
framework of upwind solvers, yielding many proposed schemes tem for t n ≤ t ≤ t n+1 ,
for approximating it. In this study, We employ the AUSM+ -up
method [11] for the inviscid fluxes. For the viscous terms Fv and
source S terms, a typical centered representation is used. η (n + 1) = (I − 0.5tAs )−1 ((I + 0.5tAs )η (n) + Bsd Fm (n))
The mesh velocity is obtained from the structural motion in (10)
response to the aerodynamic forces provided by the CFD solu- This will form part of the coupled fluid-structure (aeroelas-
tion. The structural model is described in the next section. tic) system to be elaborated below. It specifically provides the
The resulting in-house 3D RANS code has been developed time-dependent modal displacement ξ , hence the needed phys-
and validated for a variety of flow problems over a number of ical displacement of the structure so as to affect fluid flow in
years. For the validation relevant to the problem at hand will be response to the geometry variations. The mathematical system
described Section 3.1. describing the interactions between fluid and structural dynam-
ics is given below.
2.3 Structural Dynamics Equation
The finite element model for describing a structural motion 2.4 Fluid-Structural Coupling
is expressed in terms of its displacement ξ from a neutral posi- The coupling of aerodynamic and structural computations
tion (steady state in our case). In our work, we first carry out must be performed on a common geometry, while they need not
finite element analysis on a given set of nodes via MSC/Nastran be of the same mesh density or matching at the same grid points,
[12] to obtain mode shapes, Φi , i = 1, 2, · · · , Nm , Nm being the as displayed in Fig. 1 for the NASA rotor 67 blade, which is the
number of modes. Neglecting damping, the structural motion in structure that will be considered in this paper. As such, interpo-
terms of the modal displacement vector ξ = {ξi ; i = 1, · · · , Nm } lation/extrapolation procedures must be employed to accomplish
in response to the modal force Fm can be described by the mapping between them, through which the proper transfer of
relevant variables may be carried out. In our case, the structure
deformation provides a new body to the CFD process, thus af-
Mξ̈ + Kξ = Fm (7)
fecting boundary condition and the flow domain mesh. On the
other hand, the aerodynamic force needs to be transferred to the
where (M, K) are the mass and stiffness matrices of the mate- contact points for the finite element analysis. This mapping of
rial of the structure respectively. The modes on the FEM nodes grids must satisfy certain physical requirements, such as conser-
are then interpolated to every CFD node at which the aerody- vation of virtual work, and numerical requirements of accuracy
namic forces are known. This modal information facilitates the and stability.
4 Copyright
c 2014 by ASME
FIGURE 2: CFD AND CSD INTEGRATED COMPUTATION:
LOOSE COUPLING.
5 Copyright
c 2014 by ASME
advantages over other ROMs, see [7, 17] for more discussion. a system response after applying a step function,
The Volterra theory can be easily adopted as an alternative proce-
dure without having to modify the baseline full-order procedure.
σ0 , n > 0,
It is equally applicable to the time and frequency domains and σ (n) = (14)
0, n = 0.
the conversion between them is rather simple. Moreover, the for-
mulation facilitates to retain nonlinearity of the full order model
more easily than other ROMs. A small number is given to σ0 = 1.0 × 10−4 to ensure the prob-
The Volterra series, unlike the Taylor series, includes in the lem remains linear. Then, according to Eq. (13), we have the
output accumulative effects of inputs occurring at previous times. response,
The output y(t) of a continuous time-invariant system in response
to a single input u(t) for t ≥ 0 is expressed by the Volterra theory n
as: y(n) = h0 + σ0 ∑ h1 (n − k), (15)
k=0
∞ t t
y(t) = h0 + ∑ ··· hk (t − τ1 , · · · ,t − τi )u(τ1 ) · · · u(τi )d τ1 · · · d τi . And the first kernel is readily available as
i=1 0 0
(11)
where h0 is the steady-state term coincident with the initial con- 0, n = 0,
h1 (n) = (16)
dition and hi , i ≥ 1 are known as the Volterra kernels. As the time (y(n) − y(n − 1))/σ0 , n ≥ 1.
integral is discretized over a n-interval domain, a time-discrete
infinite (or truncated) Volterra series is obtained: The first equality holds because of the initial condition y(0) =
h(0).
n In what follows we show how to construct a reduced order
y(n) = h0 + ∑ h1 (n − k)u(k) model that simply bases on a relationship between the structural
k=0 motion and aerodynamic force, from the viewpoint of relating
n n
input and output data. This is easily facilitated within the state-
+ ∑ ∑ h2 (n − k1 , n − k2 )u(k1 )u(k2 ) + · · · (12)
space theory, as used in control theory. a linear state-space sys-
k1 =0 k2 =0
n n tem can be represented in the following canonical form:
+ ∑ ··· ∑ hm (n − k1 , · · · , n − km )u(k1 ) · · · u(km ) + · · ·
k1 =0 km =0
xa (n + 1) = Aa xa (n) + Ba ξ (n) (17)
Fa (n + 1) = Ca xa (n) + Da ξ (n) (18)
where y(n) is the output with the time index n referring to t n ,
u(k) is the input at preceding times k = 0, 1, 2, ..., n, and hm the
where xa (n) is the state vector at time n. The input ξ is the
mth-order Volterra kernel, m = 1, 2, · · · , ∞. For a linear system, it
structural displacement and the system output Fa denotes the non
suffices to keep only the first-order Volterra kernel, hence
dimensional generalized aerodynamic force.
To set up the above system and solve for the aeroelastic sys-
n tem under consideration, we adopt the Eigensystem Realization
y(n) = h0 + ∑ h1 (n − k)u(k) (13) Algorithm (ERA) [18]. First, we define the finite Hankel matrix
k=0
constructed using the first-order Volterra kernel h1 just described
above,
where h0 corresponds to the response with zero input, or the force
vector at steady state where there is no structural response. To ⎡ ⎤
h1 (k) h1 (k + 1) ··· h1 (k + β − 1)
capture behaviors varying with time variation, one must at least ⎢ h1 (k + 1) h1 (k + 2) ··· h1 (k + β ) ⎥
find the first-order kernel associated with the input at all other ⎢ ⎥
⎢ h1 (k + 2) h1 (k + 3) ··· h1 (k + β + 1) ⎥
times. it turns out that from the continuous system, the first ker- H(k −1) = ⎢ ⎥
⎢ .. .. .. .. ⎥
nel measures the response to an impulse applied at τ1 = 0. To ⎣ . . . . ⎦
include nonlinear effects, higher order kernels are necessary, see h1 (k + α − 1) h1 (k + α ) · · · h1 (k + α + β − 2) α ×β
Silva [8] (19)
In the present study, we make use of the first kernel to build where α and β are the sampling time shift in the row and column
our reduced order model (ROM), for which the necessary step is directions respectively; they control the order (rank) of the sys-
the definition of h1 (n), for n ≥ 0, as will be illustrated below for tem, and are set as α = 1600, and β = 50 in our study. Applying
6 Copyright
c 2014 by ASME
Singular Value Decomposition to H(0), track starts with the baseline CFD solution as the full-order will,
then builds the Volterra kernel shown in Eq. (16), which forms
H(0) = UΣV T (20) the state space system in Eq. (17). The input and output of which,
ξ and Fa , are coupled with the structural dynamics system in Eq.
(10). It is noted that Fm = q∞ Fa , with q∞ being the dynamic pres-
we find U, Σ and V , which are then used to define the matrices in sure (ρ∞ a2∞ ). These two systems combined form the ROM for the
Eq. (17): aeroelastic analysis discussed next.
where
T
EM = [IM 0M · · · 0M ]α M×M
ELT = [IL 0L · · · 0L ]β L×L (22)
0.1
singular value
7 Copyright
c 2014 by ASME
in today’s aircraft and also widely used in the turbo machinery
community for validation of CFD results, thus allowing us to
verify our proposed approach for AE analysis against previous
works, for example [23, 24, 5, 22].
In what follows, we shall first validate the CFD solution for
detailed profiles and performance map against the measured data
taken in [25]. Then the fluid-structure coupling procedure will be
described, followed by the aeroelastic calculation of the blade. A
model order reduction method based on the Volterra series is in-
troduced and applied to rotor 67 to determine the flutter behavior.
8 Copyright
c 2014 by ASME
(a) blade tip (b) leading edge (a) 10% span from shroud (b) 10% span from shroud
(a) static pressure ratio (b) total pressure ratio (c) total temperature ratio
(e) 70% span from shroud (f) 70% span from shroud
FIGURE 8: PROFILES OF STATIC PRESSURE, TOTAL
PRESSURE AND TEMPERATURE RATIOS AT AN EXIT LO-
FIGURE 9: RELATIVE MACH NUMBER CONTOURS AT
CATION WHEN THE ROTOR IS NEAR PEAK EFFICIENCY.
THREE SPANWISE SECTIONS, RESPECTIVELY 10%, 30%
AND 70% MEASURED FROM SHROUD.
model order reduction method for AE analysis.
It is noted, however, that an overestimation is found in the
static pressure ratio by the computation. This is probably caused
by several simplifications committed in our computational setup: The relative Mach contours at three spanwise sections, re-
(1) we did not assume a boundary layer profile at the inflow spectively 10%, 30% and 70% measured from the tip, are com-
boundary while in the experimental setup a solid surface is con- pared for the peak efficient condition in Fig. 9, revealing the
nected to the hub surface of the rotor, (2) the tip clearance is not nearly normal shock wave across the blade passage at the tip sec-
taken into account and instead an inviscid slip wall is assumed at tion, but subsonic or low supersonic near the root.
the casing, and (3) the hub wall is assumed adiabatic, hence pos- Finally, we plot the rotor 67 performance by the CFD solu-
sibly giving rise to a higher temperature or pressure in the layer tion in comparison with the measured values, as shown in Fig.
at the hub. This low-momentum layer at the inlet will continue 10. At the peak efficiency point, the solution gives a mass flow
to develop, growing through the rotor, resulting in a thickened rate of 33.68 kg/s, a total pressure ratio of 1.651, and an effi-
boundary layer profile, in comparison with the computed result ciency of 0.9178. The calculated results are in good agreement
which indicates a fuller profile in a thinner layer. with the data over the entire operating conditions.
9 Copyright
c 2014 by ASME
0.96 Experiment TABLE 1: MATERIAL PROPERTIES.
caculated
adibatic efficiency
10 Copyright
c 2014 by ASME
FIGURE 12: COMPARISON OF MODAL FORCE OBTAINED
BY THE FULL ORDER AND REDUCED ORDER SOLU-
TIONS.
shown in Fig. 13, the 1st structural mode eigenvalue crosses the
imaginary axis, i.e., the eigenvalue becomes positive, indicating ified in Table 1 is determined to be structurally stable under the
an amplification of structural displacement. Figure 14 displays chosen operating condition, with a high margin of safety, when
the blade displacement predicted by the 156 aeroelastic ROM at only an isolated blade is considered, this finding consistent with
the flutter condition, the third and first modes are the two most that in [5,22]. However, blade row interactions, such as the effect
dominant ones while the second mode is nearly negligible. The of upstream inlet guid vane, can induce forced vibration in rotor
dynamic pressure needed to induce flutter is q∞ = 1.455x106 Pa, blade, thus altering its flutter characteristics, see study in [19].
nearly 10 times larger than the baseline operating condition at The ROM strategy presented here can also serve as an efficient
q∞ = 1.416x105 Pa. Hence the rotor made with the material spec- and reliable way of investigating the effect of inter-blade interac-
11 Copyright
c 2014 by ASME
tions. ACKNOWLEDGMENT
Finally we remark on the primary motivation of employing This paper presents part of the multidisciplinary design anal-
ROM, while under the foremost requirement of preserving ac- ysis and optimization (MDAO) effort contributing to the research
curacy. For performing an aeroelastic analysis over a complete towards next generation transport sponsored by the subsonic
sinusoidal cycle (Eq. (23)), the full-order (CFD-CSD) model fixed wing (SFW) project, under the fundamental aeronautics
takes 10.8 hours on a Xeon(R) W3530 computer with Intel(R) program (FAP) in NASA’s Aeronautics Mission Directorate. The
Compiler compared to 0.56 seconds used by the ROM, a whop- authors are grateful for the support of SFW management team.
ping savings by over 19,200 times. This shows the tremendous
value of using the ROM when searching for the flutter bound-
ary shown in Fig. 14, or when conducting design optimization, REFERENCES
both of which will otherwise require enormous computational [1] Bendiksen, O. O., 1986. “Role of shocks in tran-
resources. sonic/supersonic compressor rotor flutter”. AIAA J., 24,
pp. 1179–1186.
[2] Menter, F., 1994. “Two-equation eddy-viscosity turbulence
st
models for engineering applications”. AIAA Journal, 32,
1 mode
nd
pp. 1598–1605.
-6 2 mode
4.0x10 [3] Liou, M. S., 2006. “A sequel to AUSM, Part II: AUSM+–up
rd
3 mode
for all speeds”. J. Comput. Phys., 214, pp. 137–170.
-6
2.0x10
displacement(m)
12 Copyright
c 2014 by ASME
[16] Raveh, D. E., Levy, Y., and Karpel, M., 2000. Aircraft
aeroelastic analysis and design using cfd-based unsteady
loads. AIAA Paper 2000–1325.
[17] Balajewicz, M., and Dowell, E., 2012. “Reduced-order
modeling of flutter and limit-cycle oscillations using the
sparse volterra series”. J. of Aircraft, 49, pp. 1803–1812.
[18] MathWorks Inc., 2012. Control system toolbox manual.
[19] Zhang, C., Ye, Z., and Liu, F., 2009. Numerical researches
on aeroelastic problem of a rotor due to igv/fan interaction.
AIAA Paper 2009–865.
[20] Bendiksen, O. O., 1990. Aeroelastic problems in turboma-
chines. AIAA Paper 1990–1157.
[21] Bartels, R., and Sayma, A., 2007. “Computational aeroe-
lastic modeling of airframes and turbomachinery: Progress
and challenges”. Phil. Trans. Royal Soc. A, 365, pp. 2469–
2499.
[22] Sadeghi, M., and Liu, F., 2005. Coupled fluid-structure
simulation for turbomachinery blade rows. AIAA Paper
2005–0018.
[23] He, L., and Denton, J. D., 1994. “Three-dimensional
time-marching inviscid and viscous solutions for unsteady
flows around vibrating blades”. J. Turbomachinery, 116,
pp. 469–476.
[24] Chuang, H. A., and Verdon, J. M., 1998. A numerical simu-
lator for three-dimensional flows throughou vibrating blade
rows. Nasa cr 1998-208511.
[25] Strazisar, A. J., et al., 1989. Laser anemometer measure-
ments in a transonic axial-flow fan rotor. NASA TP 2879.
[26] Cunnan, W. S., Stevens, S., and Urasek, D. C., 1978. De-
sign and performance of a 427-meter-per-second-tip-speed
two-stage fan having a 2.40 pressure ratio. Nasa tp-1314.
13 Copyright
c 2014 by ASME