LINFLUX-AE: A Turbomachinery Aeroelastic Code Based On A 3-D Linearized Euler Solver

Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

NASA/TM—2004-212978

LINFLUX-AE: A Turbomachinery Aeroelastic


Code Based on a 3–D Linearized Euler Solver
T.S.R. Reddy
University of Toledo, Toledo, Ohio

M.A. Bakhle, J.J. Trudell, O. Mehmed, and G.L. Stefko


Glenn Research Center, Cleveland, Ohio

May 2004
The NASA STI Program Office . . . in Profile

Since its founding, NASA has been dedicated to • CONFERENCE PUBLICATION. Collected
the advancement of aeronautics and space papers from scientific and technical
science. The NASA Scientific and Technical conferences, symposia, seminars, or other
Information (STI) Program Office plays a key part meetings sponsored or cosponsored by
in helping NASA maintain this important role. NASA.

The NASA STI Program Office is operated by • SPECIAL PUBLICATION. Scientific,


Langley Research Center, the Lead Center for technical, or historical information from
NASA’s scientific and technical information. The NASA programs, projects, and missions,
NASA STI Program Office provides access to the often concerned with subjects having
NASA STI Database, the largest collection of substantial public interest.
aeronautical and space science STI in the world.
The Program Office is also NASA’s institutional • TECHNICAL TRANSLATION. English-
mechanism for disseminating the results of its language translations of foreign scientific
research and development activities. These results and technical material pertinent to NASA’s
are published by NASA in the NASA STI Report mission.
Series, which includes the following report types:
Specialized services that complement the STI
• TECHNICAL PUBLICATION. Reports of Program Office’s diverse offerings include
completed research or a major significant creating custom thesauri, building customized
phase of research that present the results of databases, organizing and publishing research
NASA programs and include extensive data results . . . even providing videos.
or theoretical analysis. Includes compilations
of significant scientific and technical data and For more information about the NASA STI
information deemed to be of continuing Program Office, see the following:
reference value. NASA’s counterpart of peer-
reviewed formal professional papers but • Access the NASA STI Program Home Page
has less stringent limitations on manuscript at http://www.sti.nasa.gov
length and extent of graphic presentations.
• E-mail your question via the Internet to
• TECHNICAL MEMORANDUM. Scientific [email protected]
and technical findings that are preliminary or
of specialized interest, e.g., quick release • Fax your question to the NASA Access
reports, working papers, and bibliographies Help Desk at 301–621–0134
that contain minimal annotation. Does not
contain extensive analysis. • Telephone the NASA Access Help Desk at
301–621–0390
• CONTRACTOR REPORT. Scientific and
technical findings by NASA-sponsored • Write to:
contractors and grantees. NASA Access Help Desk
NASA Center for AeroSpace Information
7121 Standard Drive
Hanover, MD 21076
NASA/TM—2004-212978

LINFLUX-AE: A Turbomachinery Aeroelastic


Code Based on a 3–D Linearized Euler Solver
T.S.R. Reddy
University of Toledo, Toledo, Ohio

M.A. Bakhle, J.J. Trudell, O. Mehmed, and G.L. Stefko


Glenn Research Center, Cleveland, Ohio

National Aeronautics and


Space Administration

Glenn Research Center

May 2004
Acknowledgments

The authors thank Dr. Ed Envia, NASA Glenn Research Center, and Dr. J. Verdon, Ohio Aerospace Institute,
for many helpful suggestions on running LINFLUX. The authors also thank Mr. Tony Herrmann for
providing the finite element grid for ANSYS analysis, and Dr. R. Srivastava for providing assumed
modes. This work was supported by a grant from the NASA Glenn Research Center, Structural
Mechanics and Dynamics Branch, with funding from the Smart Efficient
Components Project. Robert Corrigan is the program manager.

Trade names or manufacturers’ names are used in this report for


identification only. This usage does not constitute an official
endorsement, either expressed or implied, by the National
Aeronautics and Space Administration.

Available from
NASA Center for Aerospace Information National Technical Information Service
7121 Standard Drive 5285 Port Royal Road
Hanover, MD 21076 Springfield, VA 22100

Available electronically at http://gltrs.grc.nasa.gov


LINFLUX-AE: A Turbomachinery Aeroelastic Code
Based on a 3–D Linearized Euler Solver

T.S.R. Reddy*
The University of Toledo
Toledo, Ohio 43606

Milind A. Bakhle, Jeffrey J. Trudell, O. Mehmed, and George L. Stefko


National Aeronautics and Space Administration
Glenn Research Center
Cleveland, Ohio 44135

Abstract
This report describes the development and validation of LINFLUX-AE, a turbomachinery aeroelastic
code based on the linearized unsteady 3–D Euler solver, LINFLUX. A helical fan with flat plate geometry
is selected as the test case for numerical validation. The steady solution required by LINFLUX is obtained
from the nonlinear Euler/Navier Stokes solver TURBO-AE. The report briefly describes the salient features
of LINFLUX and the details of the aeroelastic extension. The aeroelastic formulation is based on a modal
approach. An eigenvalue formulation is used for flutter analysis. The unsteady aerodynamic forces required
for flutter are obtained by running LINFLUX for each mode, interblade phase angle and frequency of
interest. The unsteady aerodynamic forces for forced response analysis are obtained from LINFLUX for the
prescribed excitation, interblade phase angle, and frequency. The forced response amplitude is calculated
from the modal summation of the generalized displacements. The unsteady pressures, work done per cycle,
eigenvalues and forced response amplitudes obtained from LINFLUX are compared with those obtained
from LINSUB, TURBO-AE, ASTROP2, and ANSYS.

Introduction

An overview of the aeroelastic analysis methods for turbomachines (ref. 1) shows both time and
frequency domain methods have been used to obtain unsteady aerodynamic forces and to solve the
aeroelastic equations. It was noted that time domain methods require large computational times compared to
frequency domain methods, and should only be used when nonlinearities are expected, and when the need
justifies the cost. Two approaches were used in obtaining the unsteady aerodynamic forces using frequency
domain methods. In the first approach (ref. 2) the unsteady aerodynamic equations are linearized about a
uniform steady flow, thereby neglecting the effects of airfoil shape, incidence and thickness. The unsteady
aerodynamic models developed in references 3 to 6 based on this approach were used in references 7 and 8
to study the flutter and forced response analysis of a compressor rotor and propfans. However, methods
developed by the uniform steady flow approach are restricted to shock-free flows through lightly loaded
blade rows.
In the second approach (ref. 9) the unsteady non-linear aerodynamic equations are linearized about the
non-uniform steady flow. This results in variable coefficient linear unsteady aerodynamic equations, which
include the effects of steady aerodynamic loading due to airfoil shape, thickness and angle of attack.
Following the second approach, unsteady linearized Euler aerodynamic models that include the effect of

*
NASA Resident Research Associate at NASA Glenn Research Center

NASA/TM—2004-212978 1
steady aerodynamic loading were developed in references 10 to 14. The aeroelastic code described here is
based upon the two and three-dimensional linearized Euler codes named respectively LINFLUX2D and
LINFLUX, developed in references 13 and 14 under a NASA contract. These codes were based respectively
on the non-linear Euler solver developed in references 15 to 17.
In reference 18, the unsteady aerodynamic calculations from LINFLUX2D were used with MISER
(ref. 7), an aeroelastic stability and response code based on a typical section structural model. In
reference 19, the unsteady aerodynamic calculations from LINFLUX2D were used with ASTROP2 (ref. 8),
an aeroelastic stability and response code that uses strip theory to integrate two-dimensional aerodynamic
forces on to a three dimensional structure. Flutter and forced response calculations were presented for
cascades in subsonic and transonic flow, with and without mistuning. The aeroelastic calculations with
MISER and ASTROP2 approximate the behavior of a three dimensional structure with a two dimensional
equivalent. However, a more accurate aeroelastic analysis would couple a three dimensional structural
analysis with a three dimensional aerodynamic analysis.
The primary objective of the present study is to develop a 3–D-aeroelastic code by coupling the 3–D
linearized Euler aerodynamic analysis code, LINFLUX of reference 14, with a three-dimensional structural
dynamic analysis model. A modal approach is used for flutter and forced response analysis. The resulting
code is called LINFLUX-AE to indicate the AeroElastic extension to LINFLUX code. Brief descriptions of
the formulation and method of analysis are given in the next section, followed by results and concluding
remarks.

Formulation
The aeroelastic formulation involves solutions from a linearized unsteady aerodynamic model and an
aeroelastic model coupling the unsteady aerodynamic solution with the structural dynamic solution. The
aerodynamic model and the aeroelastic formulation are described in this section.

Aerodynamic model

The unsteady aerodynamic solution is obtained from the linearized Euler solver, LINFLUX, developed
in reference 14. In this code, the linearized unsteady Euler equations are obtained by expanding the
dependent variables in the unsteady non-linear Euler equations in an asymptotic series of the form

U= U (x ) + u ( x( x , t ), t ) + higher order terms (1)

Where x is the mean position, and x is the instantaneous position. Assuming the unsteady excitations
are harmonic in time, and with the first order variable to be represented as complex valued, the above
equation can be written as

U = U (x ) + Re{u ( x ) exp(iωt )} (2)

Here, the term U (x ) is of order one and the second term is of the order ε.
Substituting the expansion of equation (2) in the non-linear unsteady Euler equations, and equating
terms of like power in ε, and neglecting terms of second and higher order in ε, nonlinear steady equations
and linear variable coefficient unsteady equations are obtained.
For harmonic blade motions with constant phase angle between adjacent blades (interblade phase
angle), the values of interblade phase angle (σ) that can occur are given as (ref. 21).

NASA/TM—2004-212978 2
σ r = 2πr N ; r = 0,1,2,……, N – 1 (3)

where N is the number of blades in the blade row. In a time domain approach with periodic boundary
conditions, the number of blade passages required to be modeled depends on the interblade phase angle, and
small phase angles may require large number of blade passages to calculate the unsteady aerodynamic
forces. However, for the linearized approach with assumed harmonic variation in time, the periodic
conditions are applied on a single extended blade passage region i.e., a region of angular pitch,

θ = 2π / N (4)

In solving the linearized unsteady equations, the dependent variables are regarded as pseudo-time
dependent. This allows solutions to be determined using conventional time-marching algorithms to
converge the steady and the complex amplitudes of the unsteady conservation variables to their steady state
values. For more details, see reference 14.
The steady aerodynamic loading required for the linearized solver, LINFLUX, is obtained from the 3–D
Euler/Navier-Stokes solver TURBO-AE, (ref. 20). The code is based on a finite volume scheme. Flux
vector splitting is used to evaluate the flux Jacobians on the left-hand side of the governing equations and
Roe's flux difference splitting is used to form a higher-order Total Variation Diminishing (TVD) scheme to
evaluate the fluxes on the right hand side. Newton sub-iterations are used at each time step to maintain
higher accuracy. A Baldwin-Lomax algebraic turbulence model is used in the code. Additional details
regarding the code are available in references 16, 17, and 20. Only the Euler option of the TURBO-AE code
is used in the present study.

Aeroelastic model

The aeroelastic equations of motion for each blade of the blade row of a turbomachine can be written as

[M ]{X }+ [K ]{X } = {P( X , t )} + {F (t )} (5)

where [M] is the mass matrix, [K] is the stiffness matrix that includes the effects of rotation (centrifugal
stiffening and softening), X} is the vector of blade deflections at the finite element grid points, {P(X,t)} is
aerodynamic force vector due to blade vibration, and F(t) is the aerodynamic force vector due to unsteady
gust or excitation independent of blade vibration. The motion dependent forces cause flutter, and motion
independent forces cause forced response (forced vibration).
A modal formulation is used to proceed further. First, a free vibration solution of the blades giving
generalized masses, natural frequencies, and mode shapes is obtained. The general vibratory motion can
then be expressed as a superposition of the contributions of the normal modes, [φ] = [φ1 , φ 2 ,......, φ NM ] , as

i = NM
{X } = [φ]T {q} = ∑ φ i qi (6)
i =1

where {q} is the generalized displacement, and NM is the number of normal modes to be used in the
analysis. Substituting equation (6) in equation (5), and post multiplying the result by [Φ]T, the equations of
motion for the sth blade can be written as

[M s ]{qs } + [K s ]{q s } = [As ]{q s } + {ADs } (7.1)

NASA/TM—2004-212978 3
where
[M s ] = [φ]T [M ][φ] ;
[K s ] = [φ]T [K ][φ] ; (7.2)
[As ] = [φ]T {P(φ), t} ;
{ADs } = [φ]T {F (t )}
[Ms] and [Ks] are generalized mass and stiffness matrices, which are diagonal, {qs}is the generalized
displacement vector, [As] is the blade vibration dependent generalized aerodynamic load matrix, and {ADs}
is the blade vibration independent generalized aerodynamic load vector. The matrices [Ms], [Ks], and [As]
are of NM x NM size, {qs} and {ADs} are of NM x 1 size.
For a tuned cascade (or rotor), in which all the blades are identical, the aeroelastic modes consist of
individual blades vibrating with equal amplitudes at a fixed interblade phase angle between adjacent blades.
Hence, the motion of the sth blade in rth interblade phase angle mode can be written as

{q s } = {qos }e iωt = {qar }e iσr s e iωt (8)

where σ r is given by equation (3).


Thus the equations of motion for the sth blade, equation (7.1) becomes

− ω 2 [M s ]{q ar }e i (ωt + σ r s ) + [K s ]{q ar }e i (ωt + σ r s ) = [ Ar ]{q ar }e i (ωt +σ r s ) + {ADr }e i (ωt +σ r s ) (9)

Since the blades are identical, the same equation is obtained for each blade. Thus, no additional information
can be obtained by assembling the equations for all the blades on the disk. Instead, equation (9) can be
solved for N different values of the interblade phase angles given by equation (3). It should be noted that for
a mistuned rotor, where the blades have different structural and/or aerodynamic properties, equation (7)
have to be assembled for all the blades in the rotor and solved, see reference 7 for more details for mistuned
formulation.
Calculation of elements of [Ms], [Ks], and [φ].—The elements of [Ms], [Ks] and [φ] can be obtained by
the free vibration analysis using a commercial code like ANSYS/ NASTRAN or by writing an analysis
program. The matrices [Ms] and [Ks] are diagonal. The elements of [Ms] and [Ks] are related as

K ii = M ii ωii 2 (10)

where ωii is the natural frequency of the ith mode. It is to be noted that usually the mode shapes are mass
normalized giving Mii = 1.0.
Calculation of elements of [Ar] and {ADr}.—The LINFLUX code is used to obtain the unsteady
pressures required for calculating the elements of [Ar] and {ADr} matrices. For flutter, the analysis is carried
out for a given frequency and interblade phase angle, ‘r’, and for a given mode of vibration. The output is
the real and imaginary components of unsteady pressures. This is repeated for N interblade phase angles
given by equation (3), for NM modes of vibration, and for a range of frequencies of interest. In the case of
forced response for an external excitation, the analysis is carried out for the unsteady excitation at the
required interblade phase angle and frequency of interest.
Calculation of elements of [Ar]: The elements of [Ar] are due to the vibration of the blade in a given
mode. Once the unsteady aerodynamic pressures are calculated using LINFLUX for a given frequency and
interblade phase angle, the elements of [Ar] are given as

NASA/TM—2004-212978 4

Arij = δi .dA prj (11.1)

where prj is the unsteady pressure due to blade vibration in jth mode for the rth interblade phase angle
obtained from LINFLUX, δi is the ith modal deflection, and dA is the elemental area. The elements of
equation (11.1) are given by

n = nnodes
Arij = ∑ δin ∗ prjn ∗ dAn (11.2)
n =1

where ‘nnodes’ is the number of nodes on the aerogrid. Equation (11.2) can be expanded as

n = nnodes
Arij = ∑ prjn (δ axin ∗ dAxn + δ ayin ∗ dAyn + δ azin ∗ dAzn ) (11.3)
n =1

where δaxi, δayi and δazi are the modal values at the node n for the ith mode, and dAx, dAy and dAz are the area
contributions to the node n.
It should be noted that the pressure in LINFLUX is normalized by (ρ * a∞2), where ρ is the reference air
density and a∞ is the reference speed of sound. In calculating area and modal values, attention should be
given to non-dimensionlization of length scaling. If L is the reference length used for scaling the geometry
and the modal values, the final Arij have to be multiplied by L**4 (L**2 for modal values and L**2 for
area). It should also be noted that the modal values (mode shape) should be divided by the reference length
when input to LINFLUX.
Calculation of elements of {ADr}: The elements of {ADr} are due to the external unsteady disturbances
coming onto the blade and may be due to acoustic, entropic, vortical or wake interactions. These
disturbances are independent of the blade vibratory motion, and therefore contribute only to the forced
response aspect of the aeroelastic problem. The LINFLUX solver calculates the unsteady pressures due to
these excitations for a given interblade phase angle, r. Let ‘pr’ be the unsteady pressure due to these
excitations on the blade surface. Then the elements of {ADr} are given by


ADri = δi .dA p r (12.1)

where δi is the ith mode shape. The elements of equation (12.1) are given by

n = nnodes
ADri = ∑ δin ∗ dAn ∗ prn (12.2)
n =1

where ‘nnodes’ is the number of nodes on the aerogrid. Equation (12.2) can be written as

n = nnodes
ADri = ∑ prn (δ axin ∗ dAxn + δ ayin ∗ dAyn + δ azin ∗ dAz n ) (12.3)
n =1

where δaxi, δayi, and δazi are the modal values at the node n for the ith mode, and dAx, dAy and dAz are the
area contributions to the node n.

NASA/TM—2004-212978 5
As mentioned before, in calculating area and modal values, attention should be given to non-
dimensionlization of length scaling. If L is the reference length used for scaling the geometry and the modal
values, the final ADri have to be multiplied by L**3 (L for modal values and L**2 for area).
It should be noted that the mode shape values are interpolated onto the aerogrid before running
LINLFUX. At each CFD grid point on the blade surface, the distance to the nearest three finite-element grid
points is calculated. Then, the modal deflections, [φ], at these three nearest neighbors are used in a bi-linear
interpolation scheme to calculate the interpolated value of the modal deflection at that CFD grid point. This
is done considering the blade’s undeflected position as the reference position. The interpolated modal
deflections δ are stored and are used by LINFLUX.
Stability calculations.
Work per cycle approach: To determine aeroelastic (flutter) stability using work per cycle approach, the
blade is forced to undergo a harmonic vibration in a normal mode with a specified frequency. The vibration
frequency is typically the natural frequency for the mode of interest, but some other frequency can also be
used. The aerodynamic forces acting on the vibrating blade and the work done by these forces on the
vibrating blade during a cycle of vibration are calculated as follows.
The blade motion is specified to be harmonic as

q (t ) = q0 sin(ωt ) (13)

where qo is the amplitude of motion and ω is the vibration frequency.


The work-per-cycle, W, done on the blade is calculated as:

W= ∫ ∫ pdA.(∂X ∂t )dt (14)


S

or,

W= ∫ ∫ pdA.δq0 ω cos(ωt )dt (15)


S

where, p = p(x,y,z,t) is the unsteady pressure on the blade surface due to blade vibration, A is the blade
surface area vector pointing into the blade surface, ∫ is the integral over the blade surface, ∫ is the
S
integral over one cycle of blade vibration. The equation for work per cycle using the linearized unsteady
aerodynamic equations is given in reference 14.
The work-per-cycle is an indicator of aeroelastic stability. The blade is dynamically unstable if the work
done on the blade during a cycle of blade vibration is positive. In other words, when W < 0 energy is
dissipated by the blade and the system is stable, but when W > 0, energy is gained by the blade and it is
unstable. Note that coupled mode flutter cannot be modeled with this approach, since the flutter mode and
frequency have to be known a priori for calculating the work.
Eigenvalue approach: An eigenvalue approach allows one to investigate coupled mode flutter. The
flutter frequencies and flutter modes are calculated rather than assumed as in aerodynamic work approach.
For a stability calculation (flutter), the motion-independent forces {ADr} are set to zero in equation (9).
Dropping the subscript s, since each blade is identical, and canceling out the exponential terms, equation (9)
can be written as

− ω2 [M ]{Y } + [K ]{Y } − [ Ar ]{Y } = {0} (16)

NASA/TM—2004-212978 6
where {Y} is same as {qar}.
Dividing equation (16) with square of an assumed frequency, ωo (for which the unsteady aerodynamic
forces are calculated)

− (ω ωo )2 [M ]{Y } + [K ] − [ Ar ]{Y } ωo 2 = {0} (17)

Rearranging, the equation (17) can be written in the standard eigenvalue problem as:

[[Pr ] − γ[Q]]{Yr } = {0} (18)

where

[Pr ] = [[K ] − [Ar ]] ωo 2 (19)

[Qr ] = [M ] (20)

and

γ = (ω ωo )2 (21)

The solution of the above eigenvalue problem results in NM complex eigenvalues of the form

ω
i =i γ = µ + iν (22)
ωo

The real part of the eigenvalue (µ ) represents the damping ratio, and the imaginary part (ν ) represents the
damped frequency; flutter occurs if µ ≥ 0 for any of the eigenvalues.
For the tuned cascade, the stability of each phase angle mode is examined separately. The interblade
phase angle is fixed at one of the values given by equation (3), and the NM x NM eigenvalue problem is
solved. The value of interblade phase angle is then changed, and the procedure is repeated for each of the N
permissible values. The critical phase angle is identified as the one, which results in the lowest flutter speed.
Forced response calculations: The equations of motion for the response of the blade are obtained from
equation (9), by setting the motion dependent unsteady aerodynamic forces, [Ar], to zero, as

− ω2 [M s ]{q ar }e i (ωt + σ r s ) + [K s ]{q ar }e i (ωt + σ r s ) = {ADr }e i (ωt + σ r s ) (23)

Since the blades are identical, the same equation is obtained for each blade. Thus, no additional
information can be obtained by assembling the equations for all the blades on the disk. Instead,
equation (23) can be solved once for the given excitation.
Dropping the subscript s, since each blade is identical, and canceling out the exponential terms,
equation (23) can be written as

− ω2 [M ]{Y } + [K ]{Y } = {ADr } (24)

where {Y} is same as {qar} and ω is the excitation frequency.


The aeroelastic response of the blades is calculated from equation (24) as

NASA/TM—2004-212978 7
{Yr } = [[K ] − ω2 [M ]] {ADr }
−1
(25)

The amplitude of the blade, {X}, is obtained by summing contributions of {Y} from all the modes.

i = NM
{X } = [φ]T {Y } = ∑ φiYi (26)
i =1

Procedure for Aeroelastic Analysis


The procedures to do flutter analysis and forced response analysis are shown in figures 1.1 and 1.2
respectively. The procedures consist of running six programs: (1) TURBO-AE, (2) INTERFACE, (3) a
structural dynamic analysis solver and the corresponding processing code, RDVIB, (4) a PRE-processor, (5)
LINFLUX, and (6) a POST-processor. The function of each program is explained in table 1. The response
analysis differs from flutter analysis in that for flutter analysis the LINFLUX code has to be run for a
selected number of modes, frequencies, and phase angles, whereas for response analysis, the LINFLUX
code has to be run only once for the given excitation.
The description of the steps to obtain an aeroelastic solution is as follows: In step 1, a steady
aerodynamic solution is obtained from TURBO-AE for the blade. The steady aerodynamic solution is
written as a database. Step 2 consists of running the INTERFACE program to rewrite the steady data base
to the format required by LINFLUX. In step 3 a vibration analysis is carried out for the blade to obtain
structural grid coordinates, generalized masses, natural frequencies, and mode shapes. The code RDVIB is
used to convert these outputs to a format required by the PRE-processor. In step 4 the PRE-processor is run
to interpolate modal values at the structural grid nodes onto the aerogrid nodes, for the required number of
modes. In step 5, the LINFLUX code is run for the mode of interest and frequency. At this stage LINFLUX
also calculates work per cycle for the mode of interest. In step 6 the POST processor is run, to calculate the
generalized forces, to calculate flutter using the eigenvalue approach, and to calculate the forced response
amplitudes.

Results
In this report, the verification of unsteady pressures, work done per cycle, eigenvalues, and forced
response amplitudes is given for a helical fan with flat plate geometry. The unsteady pressures calculated
using the LINFLUX code are compared with those obtained from the LINSUB code (ref. 2). The LINSUB
code is a two-dimensional unsteady cascade aerodynamic code based on linear unsteady aerodynamic
theory of Smith (ref. 3). The LINFLUX calculated work done per cycle is compared with those obtained
from LINSUB and TURBO-AE codes. The TURBO-AE code is a three-dimensional Euler aeroelastic code
developed in reference 20. The eigenvalue calculations are compared with those obtained from the
ASTROP2 code of reference 8. The ASTROP2 code uses strip theory to integrate two-dimensional
aerodynamic forces on to a three dimensional structure. It uses an eigenvalue approach to calculate flutter
stability. The LINFLUX calculated displacement amplitudes are compared with those obtained from the
ANSYS finite element program.

NASA/TM—2004-212978 8
Helical Fan Geometry and the Aero Grid

The helical fan configuration consists of a rotor with twisted flat plate blades enclosed in a cylindrical
duct with no tip gap. This configuration was developed by researchers to provide a relatively simple test
case for comparison with two-dimensional analyses. Note that there is no experimental data available for
this configuration.
The parameters for this configuration are such that the mid-span location corresponds to a flat plate
cascade with a stagger angle of 45°, unit gap-to-chord ratio, operating in a uniform mean flow at a Mach
number of 0.7 parallel to the blades. The rotor has 24 blades with a hub/tip ratio of 0.8. The radius at the
hub is 8.619 cm (3.395 in.) and the radius at the tip is 10.775 cm (4.244 in.). The inlet flow (axial) Mach
number is 0.495, and the rotation speed of the fan is 16,962.4 rpm, giving a relative Mach number of
approximately 0.7 at the mid span section.
A 141 by 11 by 41 grid is used for the calculations. On each blade surface, 81 points are located in the
chordwise direction, 11 points in the spanwise direction, and 41 points in blade to blade direction. The inlet
and exit boundaries are located at an axial distance of approximately 0.7 chord lengths from the blade
leading and trailing edges.

Steady Aerodynamic Solution

A steady flow solution is obtained first with the TURBO-AE solver. The input to TURBO-AE consists
of free stream Mach number, rotational speed, time step/CFL number, and back pressure ratio. The code is
run until the residuals decay to a very low order. For the present case, a time step of 0.04488 is used giving
a max CFL number of 60.0. About 1000 steps were needed to obtain the desired convergence. Reference 14
suggests that time discretization errors may result in spurious entropic / vortical boundary layer on the
surface, which may not allow for the LINFLUX code be run for unsteady solutions. It is recommended that
for blade motion and vortical excitations the total pressure loss on the blade surface should be less than 10
percent.
At the end of the execution of TURBO-AE, the steady aerodynamic solution and input data are written
to a file. The program INTERFACE is run next to convert the steady aerodynamic solution data obtained
from TURBO-AE to the form required by LINFLUX.

Validation of LINFLUX for Unsteady Aerodynamic Solutions

Unsteady solutions for prescribed blade vibrations: internally calculated modal values.—To validate
the LINFLUX code for unsteady solutions, unsteady pressures are calculated for harmonic blade vibration
in a prescribed mode. The prescribed mode shapes are rigid plunging motion perpendicular to the chord and
rigid pitching motion about the leading edge. The amplitude of vibration does not vary along the span. This
choice of mode shapes is meant to reduce the three dimensionality of the unsteady flow field for ease of
comparison with two-dimensional analyses.
LINFLUX can be run by an input choice, iflut = 1, by having these prescribed mode shapes internally
calculated. However, the purpose of this effort is to run for arbitrary mode shapes that are calculated outside
LINFLUX and to see that they are correctly read by LINFLUX. To check this, the mode shapes are
obtained by running LINFLUX with iflut = 1 and are written on unit 3. PRE is then run to interpolate the
mode shapes onto the aerogrid. But for this case this will not change the modal values at the aerogrid points
since the modal values are obtained from aerogrid to start with. So actually step 4 is not required for this
exercise. After running PRE the interpolated mode shapes are written to a file. Now LINFLUX is run with
iflut = 2, which reads the externally supplied mode shape values at aero-grid points. LINFLUX is run for
unsteady pressures and compared with those obtained with iflut = 1 and published results.

NASA/TM—2004-212978 9
Figures 2 and 3 show the unsteady pressures at mid span obtained for a reduced frequency, ω, of 1.0
(ω = ωo L/U = ωo L / a∞ M r = ω / M r , where ω = ωo L / a∞ , and L is the reference length). It should be
noted that ω is the input to LINFLUX. For the present calculations the value of speed of sound, a∞, is
assumed as 13707 in./sec; reference length is the chord which is 1.0 in., giving a value of 1527 cycles/sec
for the assumed frequency ωo. Figure 2 shows the variation of unsteady pressure with chord for plunging
motion and Fig 3 for pitching motion about leading edge. Fig 2.1 is for zero phase angle, and Fig. 2.2 is for
180° phase angle. Figures 3.1 and 3.2 show the unsteady pressure distributions for pitching motion for zero
and 90° phase angle respectively. The pressures are compared with those obtained from LINSUB code
based on Smith theory (ref. 3). To compare with LINSUB results, the unsteady pressure output from
LINFLUX is divided by square of the Mach number times the amplitude of oscillation. The pressure
distribution predictions from LINFLUX show excellent agreement with those from LINSUB. The work
done per cycle calculations, not shown here, also showed excellent agreement with LINSUB predictions.
Unsteady solutions for prescribed blade vibrations: externally supplied mode shapes.—Further
verification of iflut = 2 option carried out by creating a modal file outside the LINFLUX code. A program is
written to simulate rigid plunging motion and rigid pitching motion about mid chord for the helical fan. The
amplitude of pitching is 1 percent chord and pitching is 2° (0.03491 radians). The modal shape values thus
generated are written to a file. LINFLUX is run for these externally supplied mode shapes using iflut = 2
option. Figures 4 and 5 show the unsteady pressure difference at mid span. At this section ω = 1.0, ω = 0.7
and the Mach number M = 0.7. Figures 4.1 and 4.2 show plunging motion results for σ = 0 and 180°, and
figures 5.1 and 5.2 show results for pitching motion about mid chord for σ = 0 and 180°. They are
compared with those obtained from LINSUB. The LINFLUX calculations show excellent agreement with
LINSUB calculations for both plunging and pitching motion, since the simulated motion is close to a two
dimensional rigid plunging and pitching motion. The work done per cycle obtained from LINFLUX,
LINSUB and TURBO-AE is compared in figures 6.1 and 6.2 for pitching and plunging motions
respectively. The LINFLUX calculations show excellent agreement with LINSUB predictions. However,
TURBO-AE predictions show a slight difference from LINSUB calculations. This may be because in
TURBO-AE reflecting boundary conditions are used in obtaining the solutions. Table 2 shows the
numerical values of the work done per cycle (wpc) for both the modes from LINFLUX, TURBO-AE and
LINSUB. It should be noted that the LINSUB wpc values are multiplied with (amplitude of oscillation**2)
(M**2)*(span) with span equal to 0.849 for the present case. The excellent agreement in unsteady pressures
validates the unsteady pressure calculations with LINFLUX for externally supplied modes.
Unsteady solutions for external excitations.—To validate LINFLUX for forced response excitation,
analysis is carried out for three types of excitations: a downstream acoustic excitation, upstream acoustic
excitation, and a vortical gust. The unsteady pressure difference is compared with those obtained by
LINSUB in figures 7 to 9. It should be noted that structural mode shapes are not required for running
LINFLUX for external excitations.
Figures 7 and 8 show the unsteady pressure difference calculated for ω = 3.333 for upstream and
downstream acoustic waves for σ = 90° respectively. The results for the upstream acoustic wave show
excellent agreement with those obtained from LINSUB. The results for downstream acoustic wave show
fair agreement with those obtained from LINSUB. In the case of upstream acoustic wave the wave velocity
is same as the free stream velocity, and the grid resolution used in front of the cascade is sufficient to
resolve the wave velocity. However, the downstream acoustic wave is traveling opposite to the free stream,
and the grid aft of the cascade may require higher resolution to match the LINSUB predictions.
Figure 9.1 shows the unsteady pressures for a vortical gust at –90° phase angle and ω = 1.287, and
figure 9.2 shows the unsteady pressures for vortical gust at –180° phase angle and ω = 2.574. The unsteady
pressures are compared with those obtained from LINSUB. The figures show that the trends are correctly
predicted but the magnitudes are not predicted well. This may indicate that for the vortical gust excitations a
finer grid is required to simulate the wave lengths correctly.

NASA/TM—2004-212978 10
Structural model, mode shapes, and frequencies

The purpose of this development is to use LINFLUX for aeroelastic analysis of real blades with real
mode shapes with a mode shape file that is externally supplied to LINFLUX. For this purpose a realistic
structural dynamic analysis for structural mode shapes, generalized masses, and frequencies is required.
This will help in validating the modules PRE and POST, and streamlining the procedure for flutter and
forced response analysis of real blade geometries. For the helical fan geometry considered here there was no
structural model available. Therefore, from the available aerodynamic grid, a finite element grid was made.
The finite element grid is made exactly to match with the aerogrid by using 81 nodes along the chord and
11 nodes along the span giving 891 nodes on the surface. The blade is assumed to have thickness of 0.02
inches. The material of the blade is assumed to be titanium with a material density of 0.160 lb/in**3,
Young’s modulus is 16x10E6 lb/in**2 and the Poisson’s ratio is 0.3. The blade is rotating at 16,962 rpm
about the engine axis.
The structural dynamic analysis was performed with the ANSYS computer code. SHELL 63 elements
were used in the analysis. Dynamic analysis included the effects of spinning, and stress softening. Figures
10.1 and 10.2 show the blade modes shapes for first and second modes. The first mode is a bending mode at
822 Hz and the second mode is a torsion mode at 1882 Hz. The frequency values are tabulated in table 3 for
both rotating and non-rotating blades. For the present aeroelastic analysis the first two modes are used.

Validation of PRE

The purpose of the module PRE is to rotate the structural gird to aero grid coordinate system, and to
interpolate modal values from structural grid onto aerogrid. These interpolated modal values are required to
calculate the unsteady aerodynamic forces using LINFLUX and to calculate the generalized forces in
POST. Any non-dimensionalization of the geometry is also to be accounted for at this stage. After running
PRE the interpolated mode shapes are written to a file. The interpolated modal values are visually checked,
and also by plotting the interpolated mode shapes. Table 4 shows the modal values on the structural grid
and the interpolated values on the aerogrid at selected nodes. They show excellent agreement as expected
since the structural and aerogrid are identical for the present example.

Validation of POST

In POST, the unsteady pressures calculated by LINFLUX, and the interpolated mode shapes from PRE
are used to calculate generalized forces and displacements, eigenvalues, and response amplitudes. The
following describes how they were systematically checked.
Generalized forces.—The generalized force calculation is checked by assuming uniform pressure
on the aerogrid and assuming a modal value of 1.414 perpendicular to chord at all points. The blade face
area of the helical fan is [length * chord = (4.244 – 3.395)*1.0] = 0.849 in**2 (calculated area = 0.8491).
As the helical fan has 45° twist, the area Σdax = 0.849*0.707 = 0.62923 (0.5998) and
Σday = 0.849*0.707 = 0.62923 (0.5913) and Σdaz = 0.0 (0.09938), a quantity an order less than Σdax and
Σday. The displacement values in x, y, and z directions are 1.00, 1.00 and 0.0. Substituting these values in
equation (11.3) one obtains a value of (0.62923*1.0 + 0.62923*1.0 + 0.09938* 0.0) = 1.26 (1.19) for
generalized force. The computer calculated generalized force values show a 6.0 percent difference from
hand calculated values. This is expected since the hand calculations assumed a constant twist of 45° along
the entire span.
The generalized forces and the displacements calculations are also checked by assuming uniform
pressure on the aerogrid but using the ANSYS calculated modal values for two modes of the non-rotating
blade. Both ANSYS and POST use harmonic response analysis method to determine the steady-state

NASA/TM—2004-212978 11
response of a linear structure to loads that vary sinusoidally with time. The mode superposition method is
used to calculate the structure’s response. Table 5 shows the x, y, and z components of the displacements at
the tip leading edge (node number 811), mid chord (851) and at the trailing edge (891) obtained from both
POST and ANSYS. The excitation frequency is 400 Hz. They show excellent agreement.
Flutter.—Flutter analysis with LINFLUX is carried out using the first two modes given by ANSYS for
a frequency of 1527Hz ( ω = 0.7 and ω = 1.0). The calculated unsteady pressures are used by POST to
calculate generalized forces and eigenvalues. To validate the eigenvalue calculations, the present results are
compared with those obtained from the ASTROP2 code. As mentioned before, the ASTROP2 code uses
strip theory to integrate two-dimensional aerodynamic forces on to a three dimensional structure. The
aerodynamic forces are calculated about leading edge, using theory of reference 4. It also uses eigenvalue
approach to calculate stability.
Unsteady pressure comparison for the calculated structural mode shapes: Figures 11 and 12 show the
unsteady pressures calculated for mode1 and mode 2 respectively, for σ = 0 and 180° at ω = 0.7. They are
compared with TURBO-AE. They agree qualitatively, but show some quantitative differences. In the
previous sections it was shown that calculations from LINFLUX compared well with LINSUB predictions
for assumed modes. Since the airfoil geometry is same except for the mode shapes the LINFLUX
calculations are assumed to be accurate. The quantitative differences between TURBO-AE and LINFLUX-
AE results indicate that a more accurate TURBO-AE solution is required by optimizing the time step,
amplitude of oscillation, and number of cycles used to obtain the solution. For the present report time step,
amplitude of oscillation and number of cycles are fixed, which may affect the solution for phase angles, and
frequencies. In addition, reflecting boundary conditions are used in the TURBO-AE runs, whereas
LINFLUX uses non-reflecting boundary conditions. It should be noted that getting an optimum TURBO-
AE solution is not the objective of this report.
Work per cycle (wpc): As mentioned before, work per cycle can be calculated with LINFLUX for each
mode at a given frequency. This will give information on the stability of each mode (but not for a coupled
mode for which a coupled mode analysis has to be carried out as given in the next section). The calculated
work per cycle from LINFLUX and TURBO-AE are listed in table 6, and shown in figure 13.1 for the first
ANSYS mode and in figure 13.2 for the second ANSYS mode. All the calculations for
ω = 0.7. It can be seen that they compare very well. The discrepancies seen are due to the non-reflecting
boundary conditions used in LINFLUX and not in TURBO-AE runs.
Stability root locus: The POST program in LINFLUX-AE code is used now to calculate the stability
eigenvalues using the ANSYS calculated mode shapes. The calculations are made for 8 of the 24 possible
phase angles at ω = 0.7. ASTROP2 calculated cascade parameters and equivalent plunging and pitching
values at selected strips for this blade are given in appendix A. Table 7 shows the numerical values of
damping and frequency obtained from LINFLUX and ASTROP2. Figures 14 and 15 show the plot of the
frequency versus damping ratio for the two modes. There is only slight difference in the least stable
damping ratio. This is expected since the LINFLUX unsteady pressure calculations include the effect of
chordwise flexibility, and three dimensional flow effects. As expected, the eigenvalues indicate that the
blades are stable, which was also indicted by the work per cycle approach in the TURBO-AE and
LINFLUX programs. These results validate the routines in the post-processor for flutter prediction.
Forced response amplitudes.—The response amplitudes are checked by using the LINFLUX calculated
unsteady pressures for an upstream acoustic wave and using the ANSYS calculated modal values for two
modes for the non-rotating blades. The frequency of this excitation is 400 Hz. The unsteady pressures from
the aerogrid are interpolated on to the structural grid and an ANSYS analysis was carried out. Table 8
shows the x, y, and z components at nodes 811, 851 and 891 obtained from both POST and ANSYS. They
show good agreement for both real and imaginary values.
These results validate the POST module for generalized forces and response amplitude.

NASA/TM—2004-212978 12
Concluding Remarks
The formulation and validation of the turbomachinery aeroelastic code LINFLUX-AE are presented in
this report. The code requires execution of six programs: (1) steady aerodynamic solution, (2) interface, (3)
read structural vibration data, (4) interpolate mode shapes on to aerogrid, (5) a linearized unsteady
aerodynamic solution, (6) and flutter and response calculations.
The various modules and the interfaces between the programs have been verified for a helical fan with
flat plate airfoils. The unsteady pressures calculated from LINFLUX for both blade vibrations, and external
disturbances agreed well with those obtained from LINSUB and TURBO-AE. The flutter eigenvalues from
LINFLUX-AE are close to those obtained from ASTROP2. The response amplitudes agreed with those
obtained by ANSYS harmonic analysis for two modes.
The helical fan example validated the procedure, and all the programs required in the LINFLUX-AE
aeroelastic package for flutter and forced response. Application of the LINFLUX-AE code for real blade
geometries is intended in future.

References
1. Reddy, et al, “A review of Recent Aeroelastic Analysis Methods for Propulsion at NASA Lewis
Research Center”, NASA TP-3406, December 1993.
2. Whitehead, D.S., “Classical Two-Dimensional Methods,” Chapter III in AGARD Manual on
Aeroelasticity in Axial Flow in Turbomachines, vol. 1, Unsteady Turbomachinery Aerodynamics, (ed.
M.F. Platzer and F.O. Carta), AGARD-AG-298, March 1987.
3. Smith, S. N., “Discrete Frequency Sound Generation in Axial Flow Turbomachines,” British
Aeronautical Research Council, London, ARC R&M No. 3709, 1971.
4. Rao, B.M. and Jones, W.P., “Unsteady Airloads on a Cascade of Staggered Blades in Subsonic Flow,”
Paper No. 32, AGARD-CP-177, September 1975.
5. Adamczyk, J.J. and Goldstein, M.E., “Unsteady Flow in a Supersonic Cascade with Subsonic Leading-
Edge Locus,” AIAA Journal, vol. 16, no.12, pp. 1248–1254, December 1978.
6. Williams, M.H., “An Unsteady Lifting Surface Method for Single Rotation Propellers,” NASA CR-
4302, 1990.
7. Kaza, K.R.V. and Kielb, R.E., “Flutter and Response of a Mistuned Cascade in Incompressible Flow,”
AIAA Journal, vol. 20, no. 8, pp. 1120–1127, 1982.
8. Kaza, K.R.V., Mehmed, O., Narayanan, G.V., and Murthy, D.V., “Analytical Flutter Investigation of a
Composite Propfan Model”, Journal of Aircraft, vol. 26, no. 8, August 1989, pp. 772–780.
9. Verdon, J.M., “Linearized Unsteady Aerodynamic Theory,” Chapter II in AGARD Manual on
Aeroelasticity in Axial Flow in Turbomachines, vol. 1, Unsteady Turbomachinery Aerodynamics, (ed.
M.F. Platzer and F.O. Carta), AGARD-AG-298, March 1987.
10. Hall, K.C. Clark, W.S., and Lorence, C.B., “A Linearized Euler Analysis of Unsteady Transonic Flows
in Turbomachinery,” ASME Paper 93–GT–94, 38th IGT and Aeroengine Congress and Exposition,
Cincinnati, Ohio, May 24–27, 1993.
11. Holmes, D.G. and Chuang, H.A., “2D Linearized Harmonic Euler Flow Analysis for Flutter and Forced
Response,” Unsteady Aerodynamic, Aeroacoustics, and Aeroelasticity of Turbomachines and
Propellers, pages 213–230, Ed. Atassi, H.M, Springer-Verlag, New York, 1993.
12. Kahl, G. and Klose, A., “Computation of Time Linearized Transonic Flow in Oscillating Cascades,”
ASME Paper 93–GT–269, 38th IGT and Aeroengine Congress and Exposition, Cincinnati, Ohio, May
24–27, 1993.
13. Verdon, J.M., Montgomery, M.D. and Kousen, K.A., “Development of a Linearized Unsteady Euler
Analysis for Turbomachinery Blade Rows,” NASA CR-4677, June 1995.

NASA/TM—2004-212978 13
14. Montogomery, M.D. and Verdon, J.M., “A Three-Dimensional Linearized Unsteady Euler Analysis for
Turbomachinery Blade Rows,” NASA CR-4770, March 1997.
15. Swafford, T.W., et al, “The Evolution of NPHASE: Euler/ Navier-Stokes Computations of Unsteady
Two Dimensional Cascade Flow Fields,” AIAA Paper 94-1834, 12th Applied Aerodynamics
Conference, Colorado Springs, Colorado, June 20–23, 1994.
16. Janus, J.M., “Advanced 3-D CFD Algorithm for Turbomachinery”, Ph.D. Dissertation, Mississippi
State University, Mississippi, 1989.
17. Chen, J.P., “Unsteady Three-Dimensional Thin-Layer Navier-Stokes Solutions for Turbomachinery in
Transonic Flow,” Ph.D. Dissertation, Mississippi State University, Mississippi, 1991.
18. Reddy, T.S.R., Srivastava, R. and Mehmed, O., “Flutter and Forced Response Analysis of Cascades
Using a Two Dimensional Linearized Euler Solver,” NASA/TM—1999-209633, November 1999.
19. Reddy, T.S.R., Srivastava, R. and Mehmed, O., “ASTROP2-LE: A Mistuned Aeroelastic Analysis
System based on a Two Dimensional Linearized Euler Solver,” NASA/TM—2002-211499, May 2002.
20. Bakhle, M.A., Srivastava, R., Keith, Jr., T, Stefko, G.L., “A 3D Euler/Navier-Stokes Aeroelastic Code
for Propulsion Applications,” AIAA–97–2749, 33rd AIAA/ASME/ASME/SAE ASEE Joint Propulsion
Conference & Exhibit, July 6–9, 1997, Seattle, Washington.
21. Lane, F., “System Mode Shapes in the Flutter of Compressor Blade Rows,” Journal of the Aeronautical
Sciences, vol. 23, pp. 54–66, Jan. 1956.

Table 1.—LINFLUX-AE programs and their function.


Number Program Name Function

I TURBO-AE do steady aerodynamic analysis

convert TURBO-AE output to data format required by LINFLUX, write


II INTERFACE
on UNIT 51

(1) do free vibration analysis. At present done outside this program


through ANSYS, NASTRAN etc.
III RDVIB
(2) read vibration analysis output file and rotate the grid and mode shapes
to aero-grid coordinate system, write on UNIT 3

interpolate (calculate) mode shape values at each aero-grid point on both


IV PRE
the airfoil surfaces for all modes (UNIT 27).

calculate unsteady pressures for the given mode, frequency and interblade
V LINFLUX phase angle; repeat for all modes, frequencies and interblade phase angles
(UNIT 94).

VI POST calculate generalized forces; flutter eigenvalues and response.

NASA/TM—2004-212978 14
Table 2.—Comparison of work per cycle for externally supplied modes
(ω = 1.0, ω = 0.7, M at mid section = 0.7).
Assumed Modes
LINFLUX TURBO-AE LINSUB*
Rigid plunging
0 –0.16990–03 –0.17161–03 –0.17296–03
45 –0.22242–03 –0.20180–03 –0.22285–03
90 –0.26345–03 –0.22098–03 –0.25855–03

180 –0.46051–03 –0.41387–03 –0.46618–03


225 –0.38399–03 –0.34110–03 –0.38879–03
270 –0.28268–03 –0.22165–03 –0.28638–03
315 –0.22321–03 –0.21899–03 –0.22625–03
Rigid pitching about
midchord
0 –0.54063–03 –0.53604–03 –0.57368–03
45 –0.43992–03 –0.29169–03 –0.46181–03
90 –0.54835–03 –0.52927–03 –0.55458–03
180 –0.13668–02 –0.15622–02 –0.13974–02
225 –0.14754–02 –0.14969–02 –0.15145–02
270 –0.14312–02 –0.11703–02 –0.14766–02
315 –0.12991–02 –0.10945–02 –0.13531–02
*
LINSUB output is multiplied by amplitude**2*M**2*span of the blade (=0.849).

Table 3.—Natural frequencies for the helical fan.


Frequency, Hz Frequency, Hz
Mode
Ω = 0.0 Ω = 16962 RPM
1 364.0 822.0

2 1745.0 1882.0

3 2220.0 2517.0

4 2538.0 2992.0

5 3667.0 4046.0

6 5075.0 5165.0

7 5472.0 5803.0

8 6507.0 7530.0

9 7499.0 8176.0

10 8321.0 8877.0

NASA/TM—2004-212978 15
Table 4.—Comparison of interpolated modal values.
Mode Lefem LE Inter Mid chord Inter TE TE inter
1 (811) (851) (891)
UX –757.7949 –756.6143 –792.5469 –790.8140 –777.0029 –777.0532
UY –793.9869 –792.8906 –837.8209 –836.0615 –835.8530 –835.9058
UZ 207.3049 205.6547 155.4340 154.8228 62.2560 62.2633

Mode LEfem LE Inter Mid chord Inter TE TE inter


2
UX –1483.5699 –1452.766 –283.6709 –277.6899 1428.260 1428.260
UY –1332.3199 –1305.039 –259.3320 –253.8341 1311.449 1311.443
UZ 305.4089 299.0735 64.5855 63.5685 –136.7429 –136.7376

Table 5.—Response amplitudes for uniform pressure at 400 Hz;


(two modes of the non–rotating blade used)

Method NODE UX UY UZ

811 –0.28392 –0.29496 0.07645


POST
851 –0.29432 –0.30894 0.05721
891 –0.28647 –0.30635 0.02277

811 –0.27556 –0.28607 0.07459


ANSYS 851 –0.28451 –0.29959 0.05010
891 –0.27444 –0.29379 0.02181

Table 6.—Comparison of work per cycle (ω = 1.0, ω = 0.7).


ANSYS Modes
LINFLUX TURBO-AE
Mode 1:
0 –4.9289 –5.241
45 –6.0959 –5.635
90 –7.0073 –6.743
135 –12.601 –12.78
180 –11.986 –16.00
225 –10.343 –12.49
270 –8.0198 –8.455
315 –6.619 –7.248
Mode 2:
0 –15.053 –17.85
45 –12.226 –9.283
90 –14.383 –13.93
135 –32.38 –33.28
180 –32.66 –43.89
225 –35.55 –46.80
270 –34.821 –37.14
315 –31.902 –35.66

NASA/TM—2004-212978 16
Table 7.—Eigenvalue Comparison ω0 = 1527 Hz, ω = 0.7, a∞ = 13707 in /sec, p∞ = 11.881 psi.

Table 7.1.—σ = 0°.


Mode LINFLUX-AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.00928 0.5334 –0.01426 0.53448
2 –0.05241 1.17275 –0.08748 1.15085

Table 7.2.—σ = 45°.


Mode LINFLUX-AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.0116 0.53408 –0.01796 0.53716
2 –0.03802 1.15010 –0.06256 1.12055

Table 7.3.—σ = 90°.


Mode LINFLUX-AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.01368 0.53899 –0.02038 0.54540
2 –0.04464 1.145 –0.07099 1.10848

Table 7.4.—σ = 135°.


Mode LINFLUX–AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.02563 0.54344 –0.02389 0.56952
2 –0.08755 1.09407 –0.21854 1.0858

Table 7.5.—σ = 180°.


Mode LINFLUX–AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.02415 0.54218 –0.03735 0.56126
2 –0.09075 1.09911 –0.16117 1.00596

Table 7.6.—σ = 225°.


Mode LINFLUX–AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.02054 0.54183 –0.02917 0.55654
2 –0.10756 1.11348 –0.18212 1.03140

NASA/TM—2004-212978 17
Table 7.7.—σ = 270°.
Mode LINFLUX–AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.01561 0.53975 –0.02074 0.54952
2 –0.11459 1.13596 –0.19597 1.07218

Table 7.8.—σ = 315°.


Mode LINFLUX-AE ASTROP2
Eigenvalue Eigenvalue
real imaginary real imaginary
1 –0.01267 0.53723 –0.01692 0.54406
2 –0.11217 1.1596 –0.2046 1.11698

Table 8.—Comparison of response amplitudes for pressure due to upstream acoustic wave
(ω=3.332, σ = 90o; two modes of the non–rotating blade used at the excitation frequency = 400 Hz).
Method NODE UX UY UZ

811 (–0.34170, –0.19651) (–0.35541, –0.20379) (0.92197–01,0.52750–01)


POST 851 (–0.35584, –0.20114) (–0.37373, –0.21112) (0.69180–01,0.39128–01)
891 (–0.34941, –0.19239) (–0.37340, –0.20624) (0.27843–01,0.15262–01)

811 (–0.33145, –0.18983) (–0.34449, –0.19668) (0.89914–01,0.51209–01)


ANSYS 851 (–0.34509, –0.19318) (–0.36339, –0.20340) (0.60738–01,0.34050–01)
891 (–0.33571, –0.18356) (–0.35895, –0.19692) (0.26731–01,0.14546–01)

NASA/TM—2004-212978 18
run TURBO-AE

: steady aerodynamic solution (UNIT 52)

run INTERFACE
: processed data for LINFLUX (UNIT 51)

run PRE
:interpolated mode shapes onto aerogrid
(UNIT 27)

run LINFLUX for mode, frequency, and


interblade phase angle of interest run RDVIB
: unsteady pressures (UNIT 94) : processed free vibration data (UNIT 3)

do structural dynamic Analysis

: generalized masses, frequencies and mode


shapes
run POST
: generalized forces, flutter and response

Figure 1.1.—LINFLUX-AE flow chart for flutter.

NASA/TM—2004-212978 19
run TURBO

: steady aerodynamic solution

run INTERFACE
: processed data for LINFLUX

run LINFLUX for the excitation type,


frequency, and interblade phase angle
: unsteady pressures (UNIT 94)

run PRE
:interpolated mode shapes onto aerogrid

run POST
: generalized forces, and response
run RDVIB
: processed free vibration data

do structural dynamic Analysis

: generalized masses, frequencies and mode


shapes

Figure 1.2.—LINFLUX flow chart for forced response.

NASA/TM—2004-212978 20
5

Real

unsteady pressure difference coeff.


0

Imag.
-5

-10

lines =linflux
symbols=linsub

-15
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 2.1.—Unsteady pressures for plunging motion,


σ = 0°, helical fan, ω = 0.7, M = 0.7.

5
unsteady pressure difference coeff.

0
Real

-5

Imag.

-10

lines =linflux
symbols=linsub

-15
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 2.2.—Unsteady pressures for plunging motion,


σ = 180°, helical fan, ω = 0.7, M = 0.7.

NASA/TM—2004-212978 21
15

10

unsteady pressure difference coeff.


5 Real

0
Imag.

-5

-10
lines =linflux
symbols=linsub

-15
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 3.1.—Unsteady pressure distribution for pitching about


leading edge, σ = 0°, helical fan, ω = 0.7, M = 0.7.

20

15
unsteady pressure difference coeff.

10

5 Real

0
Imag.

-5
lines =linflux
symbols=linsub

-10
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 3.2.—Unsteady pressure distribution for pitching about


leading edge, σ = 90°, helical fan, ω = 0.7, M = 0.7.

NASA/TM—2004-212978 22
5

Lflux-pl-re
Lflux-pl-im
Lsub-pl-re
Lsub-pl-im
unsteady pressure difference

-5

-10
0 0.2 0.4 0.6 0.8 1

distance along chord

Figure 4.1.—Unsteady pressure comparison for assumed mode,


plunging, σ = 0°, ω = 0.7, M = 0.7.

Lflux-pl-re
Lflux-pl-im
Lsub-pl-re
Lsub-pl-im
unsteady pressure difference

-5

-10

-15
0 0.2 0.4 0.6 0.8 1

distance along chord

Figure 4.2.—Unsteady pressure comparison for assumed


mode, plunging, σ = 180°, ω = 0.7, M = 0.7.

NASA/TM—2004-212978 23
10

E
F E,F=Smith
5 C C,D=Linflux
unsteady pressure difference D

-5

-10
-0.2 0 0.2 0.4 0.6 0.8 1 1.2

distance along chord

Figure 5.1.—Unsteady pressure comparison, pitching


about mid-chord, σ = 0°, ω = 0.7, M = 0.7.

15

dp=-dp(real)
dp=-dp(imag) C,D=Linflux
10 C
D

5
unsteady pressure difference

-5

-10

-15

-20

-25
-0.2 0 0.2 0.4 0.6 0.8 1 1.2

distance along chord

Figure 5.2.—Unsteady pressure comparison, pitching about


mid-chord, σ = 180°, ω = 0.7, M = 0.7.

NASA/TM—2004-212978 24
0.02
LINFLUX
TURBO-AE
LINSUB

unstable
0
Work per Cycle*10**2

stable

-0.02

-0.04

-0.06
0 60 120 180 240 300 360

Interblade Phase Angle

Figure 6.1.—Work per cycle for assumed plunging mode, helical fan, ω = 0.7.

0.05
LINFLUX
TURBO-AE
LINSUB
unstable
0
stable
Work per Cycle *10**2

-0.05

-0.1

-0.15

-0.2
0 50 100 150 200 250 300 350

Interblade Phase Angle

Figure 6.2.—Work per cycle for assumed pitching mode, helical fan, ω = 0.7.

NASA/TM—2004-212978 25
8

6
unsteady pressure difference coeff.

0 Imag.

Real
-2
Real Imag.

-4

-6 lines =linflux
symbols=linsub

-8
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 7.—Unsteady pressure difference, helical fan at r/rd = 0.9, LINFLUX vs LINSUB,
upstream acoustic wave, kc = 3.332, σ = 90°.

6
unsteady pressure difference coeff.

Imag.
4

Imag.
0

-2 Real
Real
-4

-6 lines =linflux
symbols=linsub

-8
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 8.—Unsteady pressure difference, helical fan at r/rd = 0.5, LINFLUX vs LINSUB,
downstream acoustic wave, ω = 3.332, σ = 90°.

NASA/TM—2004-212978 26
12

10
unsteady pressure difference coeff.
8

4
Imag.

Real
-2 lines =linflux
symbols=linsub

-4
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 9.1.—Unsteady pressures difference, helical fan at r/rd = 0.9,


LINFLUX vs LINSUB, vortical gust, ω= 1.287, σ = –90°.

6
unsteady pressure difference coeff.

2
Imag.
Imag.
0 Real
Real
-2

-4

-6 lines =linflux
symbols=linsub

-8
0 0.2 0.4 0.6 0.8 1
distance along chord

Figure 9.2.—Unsteady pressures difference, helical fan at r/rd = 0.9,


LINFLUX vs LINSUB, vortical gust, ω = 2.574, σ = -180°.

NASA/TM—2004-212978 27
Figure 10.1.—First mode, 822 Hz.

Figure 10.2.—Second mode, 1882 Hz.

NASA/TM—2004-212978 28
0.08
Re Dp (Linflx)
Im Dp (Linflx)
Re Dp (turbo-ae)
Im Dp (turbo-ae)
Unsteady Pressure Difference Coeff. 0.06

0.04

0.02

-0.02
0 0.2 0.4 0.6 0.8 1
Distance along chord

Figure 11.1.—Unsteady pressure distribution for mode 1, helical fan, ω = 0.7, σ =0°.

0.08
Re Dp (Linflx)
Im Dp (Linflx)
RE Dp (turbo-ae)
Im Dp(turbo-ae)
0.06
Unsteady Pressure Difference Coeff.

0.04

0.02

-0.02
0 0.2 0.4 0.6 0.8 1
Distance along chord

Figure 11.2.—Unsteady pressure distribution for mode 1, helical fan, ω = 0.7, σ =180°.

NASA/TM—2004-212978 29
0.2
Re Dp (Linflx)
Im Dp(Linflx)
0.15 Re Dp (turbo-ae)
Im Dp (turbo-ae)

Unsteady Pressure Difference Coeff. 0.1

0.05

-0.05

-0.1

-0.15

-0.2
0 0.2 0.4 0.6 0.8 1
Distance along chord

Figure 12.1.—Unsteady pressure distribution for mode 2, helical fan, ω = 0.7, σ =0°.

Re Dp (Linflx)
Im Dp(Linflx)
0.2
Re Dp (turbo-ae)
Im Dp (turbo-ae)
Unsteady Pressure Difference Coeff.

0.1

-0.1

-0.2

0 0.2 0.4 0.6 0.8 1


Distance along chord

Figure 12.2.—Unsteady pressure distribution for mode 2, helical fan, ω = 0.7, σ =180°.

NASA/TM—2004-212978 30
5
LINFLUX
TURBO-AE

unstable
0
stable
Work per Cycle

-5

-10

-15
0 60 120 180 240 300 360

Interblade Phase Angle

Figure 13.1.—Work per cycle for 1st ANSYS mode, helical fan, ω = 0.7.

unstable
0
stable

-10
Work per Cycle

-20

-30

-40

LINFLUX
TURBO-AE
-50
0 60 120 180 240 300 360

Interblade Phase Angle

Figure 13.2.—Work per cycle for 2nd ANSYS mode, helical fan, ω = 0.7.

NASA/TM—2004-212978 31
0.57

0.565

0.56
frequency ratio

0.555

0.55

0.545

0.54
stable unstable

0.535
ASTROP2 o
LINFLUX σ =0

0.53
-0.04 -0.03 -0.02 -0.01 0 0.01

damping ratio

Figure 14.—Root locus plot; damping versus frequency for helical fan, ω = 0.7, mode 1.

1.2

ASTROP2
LINFLUX

o
1.15 σ = 45
frequency ratio

1.1

1.05
stable unstable

1
-0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1

damping ratio

Figure 15.—Root locus plot; damping versus frequency for helical fan, ω = 0.7, mode 2.

NASA/TM—2004-212978 32
Appendix A
Cascade Geometric and Aerodynamic Parameters
Calculations for HFAN with ASTROP2
This appendix presents the calculated values of aerodynamic and geometric parameters obtained from
the aeroelastic code ASTROP2. These are used in the calculation of flutter eigenvalues in the ASTROP2
code. The ASTROP2 calculated eigenvalues were compared with the LINFLUX-AE calculations in figures
15 and 16. The ASTROP2 code is a quasi-three dimensional aeroelastic code. It combines two-dimensional
unsteady aerodynamic loads from linear theory of Rao (ref. 4) with the vibration modes and frequencies of
a three dimensional structure. The aerodynamic loads are combined in a strip wise fashion. An equivalent
plunging, motion perpendicular to chord, and pitching motion, rotation about leading edge, are calculated at
a selected number of strips. The unsteady aerodynamic forces are calculated for these plunging and pitching
motions.
For the present analysis, nine strips are selected. The location of the strips and their coordinates are
given in table A1. The length along the blade leading edge is 0.8256 inches. The 75 percent location is at
3.8990 from the engine axis.

Table A1.—Strip location and their coordinates.


Strip number Node Number X Y Z
1 163 –0.3654 3.4449 0.9256
2 244 –0.3635 3.5274 0.9455
3 325 –0.3616 3.6100 0.9654
4 406 –0.3597 3.6925 0.9851
5 487 –0.3578 3.7751 1.0049
6 568 –0.3560 3.8576 1.0246
7 649 –0.3542 3.9402 1.0442
8 730 –0.3524 4.0228 1.0638
9 811 –0.3506 4.1054 1.0831

First step in the ASTROP2 program is to calculate the normal, tangent, and stream vector at the selected
strips. These values are given in table A2.

Table A2.1.—Tangent vector.


Strip number Node Number X Y Z
1 163 0.0236 0.99972 0.24115
2 244 0.023 0.99974 0.24116
3 325 0.023 0.99974 0.24032
4 406 0.02303 0.99973 0.23943
5 487 0.02238 0.99975 0.23931
6 568 0.02178 0.99976 0.23717
7 649 0.02178 0.99976 0.23729
8 730 0.02178 0.99976 0.23729
9 811 0.02178 0.99976 0.23181

NASA/TM—2004-212978 33
Table A2.2.—Chord line vector.
Strip number Node Number X Y Z
1 163 0.73042 0.12478 –0.67151
2 244 0.72263 0.12416 –0.67999
3 325 0.71488 0.12345 –0.68827
4 406 0.70728 0.12284 –0.69618
5 487 0.69969 0.12212 –0.70394
6 568 0.69220 0.12154 –0.71140
7 649 0.68484 0.12087 –0.71860
8 730 0.67876 0.08017 –0.72997
9 811 0.67070 0.00000 –0.74173

Table A2.3.—Normal vector.


Strip number Node Number X Y Z
1 163 –0.70141 0.19199 –0.72727
2 244 –0.70975 0.18991 –0.71959
3 325 –0.71775 0.18763 –0.71185
4 406 –0.72540 0.18538 –0.70427
5 487 –0.73299 0.18319 –0.69678
6 568 –0.74006 0.17966 –0.68939
7 649 –0.74711 0.17816 –0.68205
8 730 –0.74882 0.17696 –0.67685
9 811 –0.74115 0.17163 –0.67054

The calculated stagger, gap to chord ratio, sweep, semi chord , and strip width are given in
table A3.

Table A3.—Calculated cascade structural parameters at the strips.


PT. STN SWEEP(DEG). SEMICHD GAP/CHD STAGGER WIDTH
1 3.44490 1.35205 0.49609 0.93961 41.87744 0.16995
2 3.52740 1.31790 0.49873 0.95678 42.51256 0.08489
3 3.61000 1.31790 0.50143 0.97365 43.13840 0.08498
4 3.69250 1.31968 0.50412 0.99040 43.74040 0.08484
5 3.77510 1.28207 0.50687 1.00681 44.33916 0.08496
6 3.85760 1.24804 0.50969 1.02290 44.91832 0.08484
7 3.94020 1.24804 0.51252 1.03881 45.48364 0.08491
8 4.02280 1.24804 0.51513 1.05438 46.04495 0.08491
9 4.10540 1.24804 0.51997 1.07228 47.06836 0.08484

The pitching values are calculated by averaging the rotational values along the chord nodes of the strip.
The effect of sweep is to add contribution due to prime terms in the above table. Since the sweep angle is
very low the contribution these numbers to aeroelastic stability is negligible.

NASA/TM—2004-212978 34
Table A4.1.—Modal values for mode 1.
PT. H-VALUE A-VALUE HP-VALUE AP-VALUE HPP-VALUE APP-VALUE
1 0.70113E+02 0.51497E+02 0.86105E+03 0.28271E+03 0.47609E+04 0.59555E+03
2 0.15462E+03 0.67203E+02 0.11638E+04 0.12313E+03 0.27826E+04 0.11671E+04
3 0.26084E+03 0.75581E+02 0.13728E+04 0.54733E+02 0.20162E+04 0.47436E+02
4 0.38294E+03 0.80152E+02 0.15292E+04 0.54810E+02 0.15581E+04 0.50465E+02
5 0.51702E+03 0.84879E+02 0.16489E+04 0.48828E+02 0.13008E+04 0.15708E+03
6 0.65975E+03 0.85496E+02 0.17402E+04 0.17189E+02 0.56030E+03 0.24749E+03
7 0.80822E+03 0.87237E+02 0.17557E+04 0.23651E+01 0.73364E+02 0.00000E+00
8 0.95798E+03 0.87438E+02 0.17498E+04 0.23651E+01 0.29395E+03 0.00000E+00
9 0.11007E+04 0.87638E+02 0.16421E+04 0.23651E+01 0.18687E+04 0.00000E+00

Table A4.2.—Modal values for mode 2.


PT. H-VALUE A-VALUE HP-VALUE AP-VALUE HPP-VALUE APP-VALUE
1 –0.18718E+03 0.42099E+03 –0.20246E+04 0.41826E+04 –0.52417E+04 0.83726E+04
2 –0.36940E+03 0.79312E+03 –0.23585E+04 0.47366E+04 –0.25552E+04 0.47168E+04
3 –0.57641E+03 0.12071E+04 –0.25654E+04 0.51218E+04 –0.20088E+04 0.40234E+04
4 –0.80006E+03 0.16518E+04 –0.27344E+04 0.54204E+04 –0.19785E+04 0.26611E+04
5 –0.10376E+04 0.21171E+04 –0.28986E+04 0.55500E+04 –0.17507E+04 0.91943E+03
6 –0.12856E+04 0.25918E+04 –0.29430E+04 0.55610E+04 –0.33911E+03 –0.50781E+03
7 –0.15370E+04 0.30539E+04 –0.29464E+04 0.53440E+04 0.41650E+03 –0.67407E+04
8 –0.17783E+04 0.34785E+04 –0.27526E+04 0.47466E+04 0.34133E+04 –0.72192E+04
9 –0.19931E+04 0.38538E+04 –0.23759E+04 0.41333E+04 0.55186E+04 –0.73247E+04

The blade is rotating counter clock wise at 16,696 rpm about the engine axis. The axial Mach number is
0.495. The calculated effective Mach numbers, and reduced frequency for an assumed frequency of
1527.6 Hz is given in table A5.

REFERENCE FREQUENCY(RAD/SEC & Hz) = 9598.19434 1527.59998

Table A5.—Calculated ASTROP2 cascade analysis parameters.


PT HMACH SWEEP EFF. M SEMICHD RED. FREQ GAP/CHD STAGGER
1 0.67728 1.35205 0.67709 0.49694 0.51393 0.93961 41.87744
2 0.68483 1.31790 0.68465 0.49963 0.51101 0.95678 42.51256
3 0.69248 1.31790 0.69230 0.50239 0.50815 0.97365 43.13840
4 0.70021 1.31968 0.70003 0.50510 0.50525 0.99040 43.74040
5 0.70804 1.28207 0.70786 0.50791 0.50244 1.00681 44.33916
6 0.71594 1.24804 0.71577 0.51077 0.49969 1.02290 44.91832
7 0.72392 1.24804 0.72375 0.51364 0.49696 1.03881 45.48364
8 0.73198 1.24804 0.73181 0.51659 0.49430 1.05438 46.04495
9 0.74012 1.24804 0.73994 0.51832 0.49051 1.07228 47.06836

NASA/TM—2004-212978 35
Form Approved
REPORT DOCUMENTATION PAGE OMB No. 0704-0188
Public reporting burden for this collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources,
gathering and maintaining the data needed, and completing and reviewing the collection of information. Send comments regarding this burden estimate or any other aspect of this
collection of information, including suggestions for reducing this burden, to Washington Headquarters Services, Directorate for Information Operations and Reports, 1215 Jefferson
Davis Highway, Suite 1204, Arlington, VA 22202-4302, and to the Office of Management and Budget, Paperwork Reduction Project (0704-0188), Washington, DC 20503.
1. AGENCY USE ONLY (Leave blank) 2. REPORT DATE 3. REPORT TYPE AND DATES COVERED
May 2004 Technical Memorandum
4. TITLE AND SUBTITLE 5. FUNDING NUMBERS

LINFLUX-AE: A Turbomachinery Aeroelastic Code Based on a 3–D Linearized


Euler Solver
WBS–22–78F–30–10
6. AUTHOR(S)

T.S.R. Reddy, M.A. Bakhle, J.J. Trudell, O. Mehmed, and G.L. Stefko

7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES) 8. PERFORMING ORGANIZATION


REPORT NUMBER
National Aeronautics and Space Administration
John H. Glenn Research Center at Lewis Field E–14454
Cleveland, Ohio 44135 – 3191

9. SPONSORING/MONITORING AGENCY NAME(S) AND ADDRESS(ES) 10. SPONSORING/MONITORING


AGENCY REPORT NUMBER
National Aeronautics and Space Administration
Washington, DC 20546– 0001 NASA TM—2004-212978

11. SUPPLEMENTARY NOTES

T.S.R. Reddy, University of Toledo, Toledo, Ohio 43606, and NASA Resident Research Associate at Glenn Research
Center; and M.A. Bakhle, J.J. Trudell, O. Mehmed (retired), and G.L. Stefko, NASA Glenn Research Center. Responsible
person, T.S.R. Reddy, organization code 5930, 216–433–6083.

12a. DISTRIBUTION/AVAILABILITY STATEMENT 12b. DISTRIBUTION CODE

Unclassified - Unlimited
Subject Category: 39 Distribution: Nonstandard
Available electronically at http://gltrs.grc.nasa.gov
This publication is available from the NASA Center for AeroSpace Information, 301–621–0390.
13. ABSTRACT (Maximum 200 words)

This report describes the development and validation of LINFLUX-AE, a turbomachinery aeroelastic code based on
the linearized unsteady 3–D Euler solver, LINFLUX. A helical fan with flat plate geometry is selected as the test case
for numerical validation. The steady solution required by LINFLUX is obtained from the nonlinear Euler/Navier
Stokes solver TURBO-AE. The report briefly describes the salient features of LINFLUX and the details of the
aeroelastic extension. The aeroelastic formulation is based on a modal approach. An eigenvalue formulation is used
for flutter analysis. The unsteady aerodynamic forces required for flutter are obtained by running LINFLUX for each
mode, interblade phase angle and frequency of interest. The unsteady aerodynamic forces for forced response
analysis are obtained from LINFLUX for the prescribed excitation, interblade phase angle, and frequency. The forced
response amplitude is calculated from the modal summation of the generalized displacements. The unsteady pres-
sures, work done per cycle, eigenvalues and forced response amplitudes obtained from LINFLUX are compared with
those obtained from LINSUB, TURBO-AE, ASTROP2, and ANSYS.

14. SUBJECT TERMS 15. NUMBER OF PAGES


41
Aeroelasticity; Linearized; Euler; Flutter; Forced response; Interblade phase angle 16. PRICE CODE

17. SECURITY CLASSIFICATION 18. SECURITY CLASSIFICATION 19. SECURITY CLASSIFICATION 20. LIMITATION OF ABSTRACT
OF REPORT OF THIS PAGE OF ABSTRACT
Unclassified Unclassified Unclassified
NSN 7540-01-280-5500 Standard Form 298 (Rev. 2-89)
Prescribed by ANSI Std. Z39-18
298-102

You might also like