You are on page 1of 14

Hypersonic Aeroelasticity and Aerothermoelasticity

Jack J. McNamara1 and Peretz P. Friedmann2


1
Department of Aerospace Engineering, The Ohio State University, Columbus, OH, USA
2
Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI, USA

aerodynamics, (ii) viscous interactions between the rela-


1 Introduction 1 tively thick boundary layer, the outer inviscid flow, and
2 Fundamental Characterization of the Problem 2 shock waves, and (iii) extreme flow compression and viscous
dissipation (Anderson, 2006). These issues make accu-
3 Modeling Approaches 3
rate modeling of the unsteady aerodynamics a significant
4 Experimental Studies 7
challenge, and introduce the need to account for viscous
5 Computational Studies 8 effects and heat transfer between the fluid and structure.
6 Future Developments 12 Typically hypersonic aeroelasticity refers to aeroelastic anal-
Acknowledgements 12 ysis that neglects aerodynamic heating, while hypersonic
Nomenclature 12 aerothermoelasticity is an expanded analysis that includes
aerothermal effects. Aeroelastic phenomena are an essen-
Related Chapters 12
tial consideration in the hypersonic flow regime due to the
References 13
potential to lead to structural failures (e.g., flutter, fatigue,
Further Reading 13 buckling, etc.) and increase the complexity of interactions
between the aerodynamic, structural, control, and propulsion
systems. The latter is a concern since time varying defor-
1 INTRODUCTION mations, structural stiffness, and overall vehicle mass (as a
result of fuel burn) affect the inflow to the engine, the aerody-
Hypersonic aeroelasticity and aerothermoelasticity involves namics and flight dynamics, structural heating, and vehicle
the study of mutually coupled fluid-structure interactions that controllability (McNamara and Friedmann, 2007).
occur as a result of flexible structures operating in hyper- Historically, interest and research in hypersonic flight has
sonic flow. While there is no set Mach number that defines fluctuated, primarily due to immense technical challenges
the transition from supersonic to hypersonic flight, a flow associated with hypersonic propulsion, aerothermodynam-
is typically characterized as hypersonic starting between ics, and materials and structures. Early research centered
Mach 3.0 to Mach 5.0. The reason for a distinction is the on the NASA X-15 rocket plane in the 1960s. This era
increasing importance of certain physical characteristics with was followed by a lull in activity until the initiation of the
increasing Mach number. Specific flow characteristics that National Aerospace Plane Program in the mid-1980s. The
are relevant to hypersonic aeroelasticity and aerothermoe- goal of this program was to develop a single-stage-to-orbit
lasticity are: (i) significant nonlinearities in the unsteady (SSTO) reusable launch vehicle (RLV) that utilized a con-
ventional runway for both takeoff and landing. However,
the program was canceled due mainly to design require-
ments that exceeded the state-of-the-art. The VentureStar
program was a follow-up RLV project in the 1990s; however,

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
2 Aeroelasticity and Aeroservoelasticity

Figure 1. Generic air-breathing hypersonic vehicle configuration.

it failed during structural tests, again due to a lack of 2 FUNDAMENTAL CHARACTERIZATION


necessary technology. Despite program cancelations and OF THE PROBLEM
widespread technological challenges, there has remained a
determined focus on achieving routine and sustained hyper- Due to the multi-disciplinary nature of the problem, a
sonic flight. The primary motivations for continued research complete analysis requires inclusion of unsteady aerother-
in hypersonics is the need for a low-cost RLV and the modynamics, temperature dependent structural dynamics,
desires of defense agencies for unmanned hypersonic cruise and heat transfer analysis; as well as appropriate coupling
vehicles. mechanisms between each discipline. As shown in Fig-
The latest era in hypersonics has focused primarily on uti- ure 2, the hypersonic aeroelastic problem can be conveniently
lizing air-breathing propulsion systems in order to remove divided into two main components (Bisplinghoff, 1956;
the need for an oxidizer onboard the vehicle. This recent Culler, Crowell and McNamara, 2009): (i) an aerothermal
focus has resulted in successful air-breathing powered flights problem, and (ii) an aeroelastic problem. The aerother-
by the NASA Hyper-X experimental vehicle program and mal problem consists of computation of the aerodynamic
the University of Queensland’s HyShot program. Further- heating and resulting heat transfer between the fluid and
more, several more air-breathing flight tests are scheduled, the structure. Solution provides the temperature distribu-
at the time of this writing, for the DARPA/ONR HyFly tion, T (x, y, z, t), in the structure. There are two important
program, the DARPA/USAF FALCON program, and the issues to note. First, the aerodynamic heat flux is a func-
AFRL X-51 Single Engine Demonstrator (McNamara and tion of surface temperature, thus a mutual coupling exists
Friedmann, 2007; Ouzts, 2008). A generic geometry, based between the aerothermodynamics and heat transfer. Second,
loosely on the X-43 vehicle, is illustrated in Figure 1. As
shown, typical features of this class of vehicle are: a) a
lifting body; b) low-aspect ratio control surfaces; and c) a
fully integrated propulsion-fuselage system where the fore-
body and aftbody compress and expand the engine flow,
respectively. These features, in conjunction with the presence
of aerodynamic heating, lead to the complex aero-servo-
thermo-elastic-propulsive interactions mentioned previously.
Due to the impracticality of testing aerothermoelastically
scaled models in wind tunnels (Dugundji and Calligeros,
1962), high-fidelity, multi-disciplinary computational mod-
eling and simulation is a critical component of hypersonic
vehicle analysis and design (McNamara and Friedmann,
2007). Figure 2. Basic structure of the aerothermoelastic problem.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 3

the aerothermal problem is path dependent since it is a func- Note that there are two primary coupling mecha-
tion of both the instantaneous operating conditions and the nisms between the aerothermal and aeroelastic problems
trajectory leading to those operating conditions. Therefore, (Bisplinghoff, 1956). The first coupling mechanism involves
exact computation of the temperature distribution requires a the effect of temperature on the structural properties of the
sequential solution along each trajectory desired for analysis. aeroelastic system. Specifically, material properties (Young’s
The aeroelastic problem consists of interactions between modulus, coefficient of thermal expansion, etc.) are functions
the fluid, elastic, and inertial forces in a system where there is of temperature, while transient and spatially varying tem-
strong feedback between deformation and flow. The general perature distributions induce thermal stresses. Both of these
form of the aerothermoelastic equations of motion is effects can severely reduce the elastic stiffness of a struc-
ture. This mechanism is denoted as path ‘1’ in Figure 2. The
M(t)q̈v (t) + Cq̇v (t) + K[T (x, y, z, t)]qv (t) second coupling mechanism involves the effect of structural
deformation on modifying the aerothermodynamics. Thus,
= Qv (qv (t), q̇v (t), q̈v (t)) (1) the stiffness matrix is an implicit function of generalized
coordinates through the dependence of temperature distri-
where bution to structural deflections. This feedback coupling is
designated as path ‘2’ in Figure 2. A third form of coupling,
2  not shown here, is thermodynamic coupling between heat
ρ∞ V∞ ∂u
Q vi = Cp · dS (2) generation and elastic deformation. However, this effect is
2 S ∂qvi
generally negligible (Bisplinghoff, 1956).
and

Cp = Cp[x, y, z, qv (t), q̇v (t), q̈v (t)] (3) 3 MODELING APPROACHES
u = u(x, y, z, qv1 (t), . . . , qvn (t)) (4) As illustrated in Figure 3, there are several approaches
for aeroelastic analysis in hypersonic flow, under two pri-
The left side of equation (1) represents a summation of the mary divisions: (i) aerothermal–aeroelastic modeling (i.e.,
inertial, viscous damping, and elastic forces in the system. unsteady aerothermodynamics, heat transfer, and structural
Both the mass matrix (fuel burn) and stiffness matrix (tran- dynamics), and (ii) aerothermal aeroelastic coupling. In the
sient temperature distribution) are time dependent. The right first division, modeling approaches vary from simple approx-
side is the applied generalized aerodynamic forces, which are imate theories to the use of high-fidelity computational fluid
dependent on the structural generalized coordinates. dynamic (CFD) and finite element method (FEM) solvers.

Figure 3. Various modeling categories for hypersonic aeroelastic and aerothermoelastic analysis.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
4 Aeroelasticity and Aeroservoelasticity

However, the computational cost of such an approach makes response can be used to update the aerodynamic heating
it impractical in most cases. This has motivated the use of (Culler, Crowell and McNamara, 2009). For the second case,
either hybrid numerical-theoretical approaches, or reduced the quasi-static solution might correspond to the aeroelastic
order models (ROM). In the hybrid formulation, a CFD quasi-static trim condition for the vehicle – which even for
solver might be coupled to an approximate model for the steady-level flight is continuously changing due to softening
heat transfer and structural response. Alternatively, a sim- of the structure with increasing internal temperature.
ple approximate model for the unsteady aerothermodynamics
might be coupled to a FEM solver for the heat transfer and
structural response. For ROM approaches, the goal is to con- 3.1 Structural dynamics
struct a computationally efficient model from a high fidelity
model, such as CFD or FEM, using mathematical techniques The dynamics of hypersonic structures can be modeled
that identify and retain only the relevant physics in the fluid, similar to other flight regimes, using either analytical the-
thermal, and structural systems (Culler, Crowell and McNa- ories based on simple structural components (e.g., beams,
mara, 2009; Falkiewicz and Cesnik, 2009). plates, shells, etc.) or FEM for complicated structures. In the
There are three fundamental approaches for aerothermal– first case, the structure is typically discretized using global
aeroelastic coupling: (i) no aerothermal analysis, that is, approaches such as Rayleigh–Ritz or Galerkin methods. In
neglect of aerodynamic heating, (ii) one-way aerothermal- either global or local discretization, the equations of motion
aeroelastic coupling, and (iii) two-way aerothermal– are generally derived using Lagrange’s equations or Hamil-
aeroelastic coupling. One-way coupling neglects path ‘2’ ton’s principle and are in the form of equation (1). Commonly,
in Figure 2. This is the simplest form of aerothermoelastic the degrees of freedom in the equations of motion are further
modeling as it allows the aerothermal problem to be solved reduced using a normal mode transformation.
independently of the aeroelastic problem. Furthermore, this As shown in Figure 1, hypersonic vehicles are composed
approach has several subcategories, such as: (i) prescribing of low-aspect ratio components. Therefore, the structural
a temperature distribution in the structure, (ii) prescribing a dynamics typically resemble plates, where there are chord-
heat flux on the structure, or (iii) computing the heat flux on a wise, span-wise, and torsional deformations. Note that when
reference geometry from a model for the aerothermodynam- the aerodynamic heating is included in the analysis, the use of
ics. The first of these approximations assumes a condition of simple structural components such as equivalent beams and
thermal equilibrium in the structure. However, the combina- plates becomes questionable due to the influence of internal
tion of high surface temperatures, transient operating condi- structural layout to the heat transfer process. This, in addi-
tions, and structural deformations imply that this assumption tion to the ability to accurately model complex structures,
may be inaccurate in practice (Culler, Crowell and McNa- has resulted in FEM as the preferred approach for modeling
mara, 2009). The second approximation requires a priori hypersonic vehicle structural dynamics.
knowledge of the heat transfer response. Thus it is only valid
under the conditions/configuration that were used to measure
the input heat flux profile. Finally, the third approxima- 3.2 Approximate unsteady aerodynamics
tion accounts for transient heating and variable trajectories.
However it neglects the effect of structural deformation on The most commonly implemented unsteady aerodynamic
modifying the aerothermodynamics over the vehicle surface. theories are: piston theory, a similar theory developed by
Two-way aerothermal–aeroelastic coupling includes path Van Dyke, or Newtonian impact theory. Each of these
‘2’ in Figure 2 for the aerothermoelastic analysis. In complete approaches assume inviscid hypersonic flow and neglect real
form, this requires simultaneous solution of the aerothermal gas effects. Despite these simplifications, such approximate
and aeroelastic problems. However, this represents a signifi- tools have produced sufficiently accurate results in specific
cant computational challenge, primarily due to both large dif- cases. Hence, computational efficiency and ease of imple-
ferences in time scale between the heat transfer (minutes) and mentation make these methods attractive for preliminary
aeroelastic responses (fractions of a second) and the need for design and trend type studies of hypersonic configurations.
sequential analysis along extended vehicle trajectories (vary- Morgan, Runyan and Huckel (1958a) and McNamara and
ing from 15 min to 2 h in length). A more practical solution Friedmann (2007) provide in-depth reviews of these, as well
strategy is to implement a loose coupling procedure, where as other, methods.
either several aeroelastic time steps are taken between each Piston theory was developed by Lighthill (1953), who
aerothermal time step or a quasi-static solution for the struc- noted that at high Mach numbers the fluid dynamics are
tural deformation is used to update the aerodynamic heating. similar to the pressure on a piston moving in a one-
In the first case, a time-average of the structural dynamic dimensional channel. This results in a simple point-function
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 5

relationship between the local pressure on a surface and However, empirical evidence has demonstrated that
the normal component of fluid velocity produced by surface
  (γ−1)
γ
motion 2 (γ + 1)2 Ma2
   Cp,max =
2 w (γ + 1) w 2 γMa2 4γMa2 − 2(γ − 1)
Cp (x, y, z, t) = 2 +
M a a∞ 4 a∞  
1 − γ + 2γMa2
   × −1 (10)
(γ + 1) w 3 γ +1
+ (5)
12 a∞
is generally more accurate.
where Note that in some studies, Newtonian impact has been used
for blunt portions of a hypersonic vehicle (typically leading
∂Z(x, y, z, t) ∂Z(x, y, z, t) edge and nose regions) in conjunction with piston theory for
w= + V∞ (6) the remaining surfaces (Heeg et al., 1993).
∂t ∂x

Note that the accuracy of the piston theory pressure dimin-


ishes with increasing Mach number and surface inclination to
3.2.1 Incorporating viscous effects through an
the free stream (Lighthill, 1953; McNamara and Friedmann,
effective shape profile
2007). In terms of the hypersonic similarity parameter (κ), In hypersonic flow, viscous–inviscid interactions, where the
which is the product of Mach number and surface inclination boundary layer displaces the outer inviscid flow causing a
(Anderson, 1989), equation (5) has been shown to be rea- given body shape to appear much thicker, become signifi-
sonably accurate for approximately κ ≤ 1 (McNamara and cant (Anderson, 1989). This apparent thickness influences
Friedman, 2007). both the surface pressure distribution and the vehicle aeroe-
A similar expression to equation (5) was presented by Mor- lastic stability. In an attempt to account for this effect in
gan, Huckel and Runyan (1958b) based on work by Van Dyke aeroelastic computations, a static boundary layer displace-
(1951): ment thickness can be computed and added to the surface of
 a given body. Different approaches have utilized either semi-
 2 
2 Ma w M 4 (γ + 1) − 4β2 w empirical, compressible flat plate boundary layer relations
Cp (x, y, z, t) = 2 + a (Spain et al., 1993) or steady-state Navier–Stokes computa-
Ma β a∞ 4β4 a∞
tions to approximate the boundary layer displacement profile
(7) (McNamara and Friedmann, 2007). Results have demon-
strated that these approaches improve correlation with both
Note that this expression is derived from the nonlinear experimental and CFD-based hypersonic aeroelastic flutter
velocity potential equation for surfaces with shallow inclina- boundaries. However, in some cases there is still significant
tion angles to the flow. It is more accurate than piston theory error between approximate and experimental/CFD-based
at low supersonic Mach numbers, and becomes equivalent to flutter predictions (Spain et al., 1993; McNamara and
second-order piston theory (first two terms in equation (5)) Friedmann, 2007).
as Mach number increases.
As noted, piston theory is inaccurate for high Mach num-
ber flow and/or blunt surfaces. However, Newtonian impact 3.3 Approximate modeling of aerodynamic
theory, which is derived assuming the flow is composed of heating
a set of particles impacting a surface, has been shown to be
appropriate for such cases (Anderson, 1989). The expres- Relatively simple and efficient approximations for the aero-
sion for the quasi-steady form of Newtonian-Impact theory dynamic heating can be computed using Eckert’s reference
is given by (Morgan, Huckel and Runyan, 1958b): methods (Eckert, 1956; Anderson, 1989). The reference
  enthalpy method uses boundary layer relations from incom-
−1 w pressible flow theory with flow properties evaluated at a
Cp (x, y, z, t) = Cp,max sin 2
tan (8)
V∞ reference condition that accounts for the effects of compress-
ibility in high-speed flow. The reference temperature method
where in the original form uses the assumption of a calorically perfect gas to replace the
reference enthalpy with a reference temperature. However,
(Cp )max = 2 (9) note that at moderate hypersonic Mach numbers (Ma ≈ 8),
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
6 Aeroelasticity and Aeroservoelasticity

this simplification can result in approximately a 14% error


in the heat flux (Culler, Crowell and McNamara, 2009).
Eckert’s reference methods have been used extensively to
efficiently approximate viscous drag and convective heating
of aerospace vehicles (Anderson, 1989).

3.4 Heat transfer computation

In order to compute the heat transfer between the fluid


and the structure, the three-dimensional heat equation in
equation (11) must be solved. Typically, this is done using
numerical techniques (Tannehill, Anderson and Pletcher,
Figure 4. Basic structure of the aerothermoelastic problem when
1997), for example, finite differences or FEM. solved using coupled CFD and FEM solvers.

∂T ∂2 T ∂2 T ∂2 T 3.6 A simple example using a two


ρc = kx 2 + ky 2 + k z 2 (11)
∂t ∂x ∂y ∂z degree-of-freedom, double-wedge
typical section

In order to illustrate the problem with a simple example,


3.5 Coupled CFD-FEM for high-fidelity analysis consider the double-wedge airfoil shown in Figure 5a which

As computational power has increased, coupled CFD-FEM


aeroelastic analysis has emerged as a viable alternative,
and significant improvement in many cases, to experimen-
tal testing and approximate theories. In terms of an unheated
CFD-FEM aeroelastic analysis in hypersonic flow, there are
only practical differences compared to other flow regimes;
for example, time step and grid construction (McNamara and
Friedmann, 2007). Specifically, there are high spatial gradi-
ents in hypersonic flow associated with shocks and expansion
fans. This requires a high grid density in the fluid domain near
surface gradients. Furthermore, the boundary layers in hyper-
sonic flow are relatively thick. This enables a less dense grid
near the surface of the wall for boundary layer capturing.
These two issues combined result in similar cell distribu-
tion for Euler and Navier–Stokes flow analysis. For temporal
accuracy, the high speed flow implies that smaller time steps
may be needed compared to other flow regimes to capture
important unsteady aerodynamic effects.
For a coupled CFD-FEM aerothermoelastic analysis, the
aerodynamic heating must be extracted from a CFD solu-
tion to the Navier–Stokes equations, while the FEM solver
is used to perform the heat transfer analysis (McNamara and
Friedmann, 2007; McNamara et al., 2008). In this case, the
separation of aerothermal and aeroelastic problems becomes
fundamentally different, as illustrated by comparing Fig-
ures 2 and 4. In the coupled CFD-FEM approach, the problem
does not divide naturally into distinct aerothermal and aeroe-
lastic problems. Thus, in addition to significantly increased Figure 5. Example of aeroelastic behavior in hypersonic flow. (a)
Simple two degree-of-freedom, double-wedge airfoil; (b) compar-
computational requirements, developing an accurate and ision of the aeroelastic behavior of a double-wedge airfoil using
efficient coupling process presents a challenge. several different orders of piston theory aerodynamics.
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 7

has both pitch and plunge degree-of-freedom. The equations Table 1. Parameters describing the double-
of motion for this system, assuming small deflections and wedge airfoil.
no structural damping, can be obtained using Lagrange’s
equations Parameter Units Value

mḧ + Sα α̈ + Kh h = −L b m 1.175
τ — 0.0336
Sα ḧ + Iα α̈ + Kα α = MEA (12) m kg m−1 94.2
rα — 0.484
Using the small displacement assumption, the airfoil dis- ωh Hz 13.4
placement vector is given by ωα Hz 37.6
xα — 0.2
u(x, t) = −[h(t) + (x − ba)α(t)] k̂ (13)
  
The aerodynamic generalized forces can be computed 1 1 α̇ 3
MEA3 = − p∞ γ(γ + 1)Ma b 3 2
b
for piston theory aerodynamics from equations (2), (5), (6), 3 5 V∞


and (13) ḣ α̇
−a − ba +α
V∞ V∞
L = L1 + L2 + L3 (14) 

2
MEA = MEA1 + MEA2 + MEA3 ḣ α̇
(15) × − ba + α + 3τ 2  +
V∞ V∞
where 
2

α̇  ḣ α̇ α̇
b − ba + α + τ 2 − ba
ḣ α̇ V∞ V∞ V∞ V∞
L1 = 4p∞ γMa b − ba +α (16a)
V∞ V∞

ḣ α̇
× − ba +α
α̇ V∞ V∞
L2 = −p∞ γ(γ + 1)Ma2 b2 τ (16b)
V∞ (17c)


The flutter boundary for airfoil (unheated) is shown in
1 ḣ α̇ Figure 5b as a function of elastic axis offset. The parameters
L3 = p∞ γ(γ + 1)Ma b
3
− ba +α
3 V∞ V∞ used to generate these results are provided in Table 1. It is

2 
 2  evident from the divergence of the first-order piston theory
ḣ α̇ α̇
× − ba + α + 3τ 2 + b  (which is linear) prediction from the second-and third-order
V∞ V∞ V∞  results, that nonlinear aerodynamic effects play a critical role
in hypersonic aeroelasticity.
(16c)

4 EXPERIMENTAL STUDIES
and
Most experimental studies on hypersonic aeroelasticity and
    aerothermoelasticity were conducted in the 1950s, with lim-
ḣ b α̇
MEA1 = 4p∞ γMa b2 a − + ba +α ited work also conducted during the NASP program. The
V∞ 3a V∞ earliest research focused on flutter boundary identification
(17a) of elastically and dynamically scaled prototype configura-
tions, such as the X-15. However, in these cases flutter
was not observed due to high model stiffness and limited
 
ḣ α̇ wind tunnel operating conditions. Theoretical predictions
MEA2 = p∞ γ(γ + 1)Ma2 b2 τ − 2ba +α indicated the flutter boundaries were well outside the achiev-
V∞ V∞
able tunnel operating conditions. In subsequent studies, the
(17b) focus shifted to trend investigations, in order to expand

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
8 Aeroelasticity and Aeroservoelasticity

fundamental knowledge and validate theoretical approaches. 5 COMPUTATIONAL STUDIES


General observations from this early work are
The majority of recent work has utilized computational
1. Increasing airfoil thickness and angle of attack is modeling and simulation to investigate aeroelastic and
destablizing. aerothermoelastic behavior in hypersonic flow. The sam-
2. Increasing airfoil bluntness is stabilizing for small lead- ple results discussed here provide a general representation
ing edge radii (1% of chord), and destabilizing for larger of current understanding. An in-depth review is provided
leading edge radii (> 3% of chord). by McNamara and Friedmann (2007). Note that the exam-
3. The effect of leading edge sweep on flutter is inconclu- ples presented here primarily focus on dynamic stability
sive. prediction. However, the need to account for aero-servo-
4. Slab airfoil sections are more stable than double-wedge thermo-elastic-propulsive interactions in hypersonic vehicles
airfoils. implies that aeroelastic and aerothermoelastic response pre-
5. Viscous effects are destabilizing, presumably due to diction in hypersonic flow will play an increasing role in the
inviscid–viscous interactions from the boundary layer near future (McNamara and Friedmann, 2007).
displacement effect.

For a more in-depth review on past experimental inves- 5.1 Unheated analysis
tigations of aeroelastic and aerothermoelastic behavior in
hypersonic flow, refer to McNamara and Friedmann (2007). The majority of hypersonic aeroelastic studies have focused
on unheated behavior. Representative results are presented in
Figures 6 and 7 for a generic hypersonic vehicle resembling
4.1 Aerothermoelastic scaling the X-33 RLV, and Figure 8 for a generic hypersonic con-
trol surface based on the Lockheed F-104 low-aspect ratio
The development of aerothermoelastic similarity laws for wing. The analysis used to compute the results in Figure 6
experimental testing of hypersonic vehicles has been lim-
ited. The bulk of literature published on the subject was in the
180
late 1950s and early 1960s. A in-depth treatment of aerother-
moelastic similarity is provided by Dugundji and Calligeros
(1962), however, it was primarily intended for Mach num- 160 Empty

bers less than 3.5, and temperatures less than 1000◦ F. Despite 10% fuel
50% fuel
these limitations, Dugundji and Calligeros provide important 140 100% fuel
1/2 stiff/empty
insight into aerothermoelastic similarity for hypersonic flow. 1/2 stiff/100% fuel
Specifically, for the general aerothermoelastic problem, one 120
is practically restricted to full scale testing. However, for spe-
cialized cases (e.g., wing structures, thin solid plates, panel
100
flutter) the similarity laws are less restrictive and scaled model
Mc

testing is possible. Furthermore, in cases where scaled test-


ing is difficult to impossible, two alternatives are proposed by 80
Dugundji and Calligeros (1962): (i) incomplete aerothermoe-
lastic testing, and (ii) restricted purposed models. The first 60
method relies on a priori estimation of the aerodynamic pres-
sures and/or heating rates that are applied artificially to the 40
model. The latter is for cases where certain couplings can be
neglected between the aerodynamic pressure, aerodynamic
20
heating, heat conduction and stress-deflection phenomena.
Finally, note that the development of aerothermoelastically
scaled models becomes increasingly difficult as Mach num- 0
0 20 40 60 80 100 120 140
ber increases. This is due to the need for additional similarity Altitude (1000 ft)
laws that account for material and gas property variations
with temperature, radiation effects, real-gas effects, viscous Figure 6. The effect of fuel burn and vehicle stiffness variations on
interactions, and so on. the flutter boundaries of a generic hypersonic vehicle.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 9

aries are quite sensitive to variations in vehicle stiffness as


well as reductions in vehicle mass due to fuel burn. In Fig-
ures 7 and 8, two different grid resolutions were considered
during computation of the flutter boundaries. As the Mach
number was increased, the coarser grids diverged from the
finer grid predictions, illustrating the importance of increas-
ing fluid grid density near the vehicle surface as Mach number
increases and the shock moves closer to the body. Figure 8
illustrates that the importance of viscosity is dependent on
the operating conditions. At higher Mach numbers and alti-
tudes, the inclusion of viscosity did not significantly alter
the flutter boundary predictions. However, at lower altitudes
and Mach numbers, the inclusion of viscosity did affect the
flutter boundary prediction. Finally, note that the differences
between the Euler and third-order piston theory flutter Mach
numbers for the wing was only 5–8%, while the differences
grew to over 25% when the complete vehicle was considered.
This is due to increased three-dimensional flow effects on the
Figure 7. Flutter boundary of a generic hypersonic vehicle, calcu- complete vehicle, which cannot be captured by piston theory
lated using third-order piston theory and Euler aerodynamics.
aerodynamics.

consisted of an equivalent plate representation for the


structure in conjunction with first order piston theory aero- 5.2 Aerothermoelastic studies
dynamics (Friedmann et. al., 2004). The results presented in
Figures 7 and 8 were computed using a refined analysis where Despite the importance of aerodynamic heating on aeroelastic
FEM was used to model the structure, and either third order behavior in hypersonic flow, the complexity of the problem
piston theory or CFD was used for the unsteady generalized has limited aerothermoelastic research. Three sample studies
aerodynamic forces (McNamara et al., 2005, 2008). are presented in Figures 9–11. Using the categories defined
From each of these results it is clear that for a given in Figure 3, these coincide with:
altitude, hypersonic vehicles will flutter at unrealistically
high Mach numbers when aerodynamic heating is neglected. 1. Modeling Category II(ii), Coupling Category I(i) applied
Furthermore, from Figure 6 it is clear that the flutter bound- to the NASP;
2. Modeling Category II(ii) and III, Coupling Category
II(iii) applied to a low-aspect ratio wing; and
3. Modeling Category I, Coupling Categories II(iii), and
III(i–iii) applied to a surface panel.

The results presented in Figure 9 were computed using


two different structural configurations; a titanium–aluminide
(base-line configuration) and a carbon–carbon configuration
(Doggett et. al., 1991). The temperature distributions were set
to the surface temperature radiation equilibrium conditions.
From these results it is apparent that aerodynamic heating
reduces the flutter margin. In Figure 10, the aerothermoelas-
tic behavior of a low-aspect ratio wing was computed along
a representative trajectory resembling a proposed trajectory
for the DOD/USAF FALCON hypersonic cruise vehicle
(McNamara et al., 2008). As shown, the trajectory cycles
between atmospheric and near-space operating conditions in
Figure 8. Flutter boundary of a low-aspect ratio wing, calculated
using third-order piston theory, Euler, and Navier–Stokes aerody- order to mitigate heating of the structure. Aerodynamic heat-
namics. ing along the trajectory resulted in up to 30% reductions in

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
10 Aeroelasticity and Aeroservoelasticity

Figure 9. Aerothermoelastic flutter boundary of a generic hypersonic vehicle.

Figure 10. Aerothermoelastic flutter margin of a low-aspect ratio wing along a representative hypersonic trajectory.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 11

the modal frequencies of the wing, up to 40% reduction in the


aerothermoelastic flutter margin, and thermal buckling in sev-
eral cases prior to the completion of the mission. Note that the
aerothermoelastic boundary shown represents a flutter mar-
gin, rather than a true flutter boundary, since it was computed
by holding the Mach number and free vibration frequencies
constant at several points on the trajectory, and increasing
the dynamic pressure until the aeroelastic response became
unstable. Despite significant decreases in modal frequencies
and flutter margin, the minimum (pre-buckled) ratio of virtual
flutter dynamic pressure to freestream dynamic pressure was
O(10); suggesting the wing would remain dynamically stable
up to the buckling condition for the considered trajectories.
This is due to vehicle operation at high altitudes.
Finally, Figure 11 provides a comparison of the flutter
boundaries as a function of Mach number and flight time
for each of the aerothermoelastic modeling cases described
in Table 2 for a surface panel (Culler, Crowell, and McNa-
mara, 2009). Note that only steady, level trajectories were
considered, at each of the Mach numbers specified. Sev-
eral important conclusions can be drawn from these results.
First, there is no apparent threshold Mach number below
which surface panel response is aerothermoelastically sta-
ble independent of flight time. Furthermore, the neglect of
temperature dependent material properties (B-1 and B-2 )can
lead to a significant over prediction of aerothermoelastic sta-
bility. Another important result shown in Figure 11(b) is a
difference in flutter boundary prediction between one-way
aerothermal–aeroelastic coupling (C-1, D-1, and E-1) and
two-way coupling (C-2, D-2, and E-2). For the Mach 7.5 tra-
jectory, the one-way coupled cases predicted a 24% greater
Figure 11. Aerothermoelastic flutter boundary predictions for a flight time to the onset of flutter than the two-way coupled
hypersonic surface panel. (a) Comparison of flutter boundaries for cases. Two important conclusions can be drawn from this
the aerothermoelstic modeling cases; (b) Comparison of flight time
to the onset of flutter at M trajectories for the aerothermoelastic result. First, one-way coupling underpredicts the rate of mate-
modeling cases that include material property degradation. rial property degradation compared to two-way coupling.
Second, the discrepancies between one-way and two-way

Table 2. Aerothermoelastic modeling cases.

Temperature Aerothermal- Aeroheating Structural


dependent aeroelastic panel model
Case properties coupling deformation formulation

B-1 None 1-way Undeformed Quasi-static


B-2 None 2-way Instantaneous Quasi-static
C-1 E(T ), α(T ) 1-way Undeformed Quasi-static
C-2 E(T ), α(T ) 2-way Instantaneous Quasi-static
D-1 E(T ), α(T ) 1-way Undeformed Dynamic
D-2 E(T ), α(T ) 2-way Instantaneous Dynamic
E-1 E(T ), α(T ) 1-way Undeformed Dynamic
E-2 E(T ), α(T ) 2-way Time-averaged Dynamic

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
12 Aeroelasticity and Aeroservoelasticity

coupling increase with decreasing Mach number since the K stiffness matrix
panel has the largest unheated flutter margin at low Mach Kα , Kh spring constants in pitch and plunge, respectively
numbers. Thus, longer operation is required to reduce the kx , ky , kz thermal conductivity
panel stiffness to the point of flutter; indicating that errors L lift per unit span
introduced by one-way coupling build with time. This result M mass matrix
implies that two-way coupling is required to accurately pre- m mass per unit span
dict aerothermoelastic responses during vehicle operation on Ma freestream mach number
extended trajectories. MEA moment per unit span about the elastic axis
p∞ freestream pressure
Qv generalized force
6 FUTURE DEVELOPMENTS qv generalized displacement
S surface area
While significant progress has been made in the area, a num-
Sα static mass moment per unit span about elastic axis
ber of unanswered question indicate that the state-of-the-art
rα nondimensional radius of gyration
is not advanced to the point of providing reliable aerother-
T temperature distribution
moelastic analysis of hypersonic vehicles and components.
t time
For instance what role do currently unconsidered aerother-
u displacement vector
modynamic effects play in the aerothermoelastic behavior,
V∞ freestream velocity
and vice versa (e.g., hypersonic boundary layer transition,
w velocity component perpendicular to freestream
shock-turbulent boundary layer interactions, wall catalycity,
x, y, z cartesian coordinates in the freestream, span,
ablation, non-equilibrium real gas effects, turbulence induced
and vertical directions, respectively
fluctuating pressures, etc.)? How sensitive is the aerother-
xα nondimensional center of gravity offset from
moelastic behavior to the mechanical and thermal boundary
elastic axis
conditions? What role will damage accumulation have in
Z position of structural surface
the aerothermoelastic behavior during a progressive response
α airfoil
 Pitch displacement about the elastic axis
analysis? How much validation can ground-based experimen-
β Ma2 − 1
tal facilities provide?
γ ratio of specific heats
The answers to these questions require several years of
κ hypersonic similarity parameter
both basic exploration and targeted study of the problem,
ρ∞ Freestream density
as well as significant progress in numerical methodolo-
ρ structural density
gies for coupled analysis; non-deterministic approaches
τ thickness ratio
for uncertainty propagation and characterization; structural
ωh , ωα frequency associated with pitch and plunge
mechanics; and hypersonic aerothermodynamics.
stiffness, respectively

ACKNOWLEDGMENTS
RELATED CHAPTERS
The authors acknowledge the support of the NASA Funda-
mental Aeronautics Program under Cooperative Agreement Introduction to Hypersonic Flow
NNX08AB32A. Analytical Foundation of Hypersonic Flow
Fundamentals of Heat Transfer and Thermal Physics
Engineering Analysis of Heat Transfer
NOMENCLATURE Introduction to Computational Fluid Dynamics
The Scramjet Engine: Basic and Combined Cycles
a nondimensional elastic axis offset Finite Element Methods of Structural Analysis
a∞ freestream speed of sound Elements of Structural Dynamics
b semi-chord Dynamic Response Computations
C damping matrix The Evolution of Analytic and Computational Methods for
c specific heat of the structural material Fixed-Wing Flight Vehicle Aeroelasticity
Cp pressure coefficient Static and Dynamic Aeroelasticity
h airfoil vertical displacement at elastic axis Panel Flutter
Iα mass moment of inertia about the elastic axis Thermal Analysis

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Hypersonic Aeroelasticity and Aerothermoelasticity 13

REFERENCES McNamara, J.J., Friedmann, P.P., Powell, K.G., Thuruthimattam,


B.J. and Bartels, R.E. (2008) Aeroelastic and aerothermoelastic
behavior in hypersonic flow. AIAA J., 46(10), 2591–2610.
Anderson, J.D. (2006) Hypersonic and High Temperature Gas
Dynamics, 2nd edn AIAA. Morgan, H.G., Runyan, H.L. and Huckel, V. (1958a) Theoretical
considerations of flutter at high mach numbers. J. Aeronaut. Sci.
Bisplinghoff, R.L. (1956) Some structural and aeroelastic consid-
25(6), 371–381.
erations of high speed flight. J. Aerosp. Sci., 23(4), 289–329.
Morgan, H.G., Runyan, H.L. and Huckel, V. (1958b) Procedure
Culler, A.J., Crowell, A.C. and McNamara, J.J. (2009) Studies of
for calculating flutter at high supersonic speed including cam-
fluid-structural coupling for aerothermoelasticity in hypersonic
ber deflections, and comparison with experimental results. NACA
Flow. 50th AIAA/ASME/ASCE/AHS/ASC Structures, Struc-
Technical Note 4335.
tural Dynamics, and Materials Conference, May 2009, Palm
Springs, CA. AIAA-2009-2364. Ouzts, P.J. (2008) The joint technology office on hypersonics. 15th
AIAA International Space Planes and Hypersonic Systems and
Doggett, R.V., Ricketts, R.H., Noll, T.E. and Malone, J.B.
Technologies Conference, April 2008, Dayton, OH. AIAA-2008-
(1991) NASP aeroservothermoelasticity studies. NASA Technical
2576.
Memorandum 104058.
Spain, C., Zeiler, T.A., Bullock, E. and Hogde, J.S. (1993) A flut-
Dugundji, J. and Calligeros, J. (1962) Similarity laws for aerother-
ter investigation of all-moveable NASP-like wings at hypersonic
moelastic testing. J. Aerosp. Sci. 29(3), 935–950.
speeds. 34th AIAA/ASME/ASCE/AHS/ASC Structures, Struc-
Eckert, E.R.G. (1956) Engineering relations for heat transfer and tural Dynamics and Materials Conference, April 1993, La Jolla,
friction in high-velocity laminar and turbulent boundary layer CA. AIAA-1993-1315.
flow over surfaces with constant pressure and temperature. Trans.
Tannehill, J.C., Anderson, D.A. and Pletcher, R.H. (1997) Com-
ASME 78(6), 1273–1283.
putational Fluid Mechanics and Heat Transfer, 2nd edn, Taylor
Falkiewicz, N.J. and Cesnik, C.E.S. (2009) A Reduced order Francis, New York.
modeling framework for integrated thermo-elastic analysis of
Van Dyke, M.D. (1951) A Study of second-order supersonic flow
hypersonic vehicles. 50th AIAA/ASME/ASCE/AHS/ASC Struc-
theory. NACA Technical Report 1081.
tures, Structural Dynamics, and Materials Conference, May 2009,
Palm Springs, CA. AIAA-2009-2308.
Friedmann, P.P., McNamara, J.J., Thuruthimattam, B.J. and Nydick,
I. (2004) Aeroelastic analysis of hypersonic vehicles. J. Fluids FURTHER READING
Struc., 19(5), 681–712.
Heeg, J., Zeiler, T., Pototzky, A., Spain, C. and Engelund, W. Bertin, J.J. and Cummings, R.M. (2003) Fifty years of hypersonics:
(1993) Aerothermoelastic analysis of a NASP demonstrator where we’ve been and where we’re going. Prog. Aerosp. Sci. 39,
model. 34th AIAA/ASME/ASCE/AHS/ASC Structures, Struc- 511–536.
tural Dynamics and Materials Conference, April 1993, La Jolla, Bisplinghoff, R.L, Ashley, H. and Halfman, R.L. (1955) Aeroelas-
CA. AIAA-1993-1366. ticity. Addison-Wesley, Reading, MA.
Lighthill, M.J. (1953) Oscillating airfoils at high mach numbers. J. Bisplinghoff, R.L and Dugundji, J. (1958) Influience of Aero-
Aeronau. Sci., 20(6), 402–406. dynamic Heating on Aeroelastic Phenomena. Chapter 14 of
McNamara, J.J., Friedmann, P.P., Powell, K.G., Thuruthimattam, Agardograph No. 28, High Temperature Effects in Aircraft Struc-
B.J. and Bartels, R.E. (2005) Three-dimensional aeroelas- tures (ed. N.J. Hoff) Pergamon Press, Oxford, pp. 288–312.
tic and aerothermoelastic behavior in hypersonic flow. 46th Bisplinghoff, R.L and Ashley, H. (1962) Principles of Aeroelastic-
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynam- ity. John Wiley & Sons, New York.
ics and Materials Conference, April 2005, Austin, TX.
AIAA-2005-2175. Garrick, I.E. (1963) A survey of aerothermoelasticity. Aerosp. Eng.,
22(1), 140–147.
McNamara, J.J. and Friedmann, P.P. (2007) Aeroelastic and
aerothermoelastic analysis of hypersonic vehicles: Current status Rogers, M. (1958) Aerothermoelasticity. Aerospace Engineering,
and future trends. 48th AIAA/ASME/ASCE/AHS/ASC Struc- 17(10) 34–43, 64.
tures, Structural Dynamics and Materials Conference, April Thornton, E. (1996) Thermal Structures for Aerospace Applications.
2007, Honolulu, HI. AIAA-2007-2013. AIAA, Resten VA.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae153

You might also like