• No results found

On an Efficient Method fo Time-Domain Computational Aeroelasticity

N/A
N/A
Protected

Academic year: 2022

Share "On an Efficient Method fo Time-Domain Computational Aeroelasticity"

Copied!
31
0
0

Loading.... (view fulltext now)

Full text

(1)

Time-Domain Computational Aeroelasticity

David Eller

Aeronautical and Vehicle Engineering Royal Institute of Technology SE-100 44 Stockholm, Sweden

TRITA/AVE 2006:1

ISBN 91-7178-243-5

(2)

As our circle of knowledge expands,

so does the circumference of darkness surrounding it.

Albert Einstein

TRITA-AVE 2006:1 KTH Farkost & Flyg

ISSN 1651-7660 Teknikringen 8

ISRN KTH/AVE/DA–06:01–SE 100 44 Stockholm

ISBN 91-7178-243-5

Akademisk avhandling som med tillst˚and av Kungliga Tekniska H¨ ogskolan tisdagen den 24 januari 2006 kl 13, Kollegiesalen F3, Lindstedtsv¨agen 26, Stock- holm, f¨ or teknisk doktorsgrads vinnande framl¨agges till offentlig granskning av David Eller.

Typsatt i L A TEX

Tryck: Universitetsservice US AB, Stockholm 2005 c

David Eller 2005

(3)

Preface

The work documented in this thesis was performed between October 2001 and November 2005 at the Department of Aeronautical and Vehicle Engineering at the Royal Institute of Technology in Stockholm. The financial support provided by the Swedish Research Council (Vetenskapsr˚adet) and the National Program for Aeronautics Research (NFFP) is gratefully acknowledged. Part of the work was also conducted within the European Union project Active Aeroelastic Aircraft Structures.

I would like to express my gratitude to my supervisor, Professor Ulf Ringertz, for his valuable advice not limited to technical matters. Furthermore, I am grate- ful to my co-authors Martin Carlsson for thoroughness, and Sebastian Heinze for ingenuity, my colleagues Dan Borglund (”Piece of cake!”) for optimistic com- ments when confronted with doubt, Marianne Jacobsen and Carin Cronander for inspiration and well-considered comments.

Stockholm, December 2005

David Eller

(4)
(5)

Abstract

The present thesis summarizes work on developing a method for unsteady aero- dynamic analysis primarily for aeroelastic simulations. In contrast to widely used prediction tools based on frequency-domain representations, the current approach aims to provide a time-domain simulation capability which can be readily integrated with possibly nonlinear structural and control system models.

Further, due to the potential flow model underlying the computational method, and the solution algorithm based on an efficient boundary element formulation, the computational effort for the solution is moderate, allowing time-dependent simulations of complex configurations.

The computational method is applied to simulate a number of wind-tunnel experiments involving highly flexible models. Two of the experiments are uti- lized to verify the method and to ascertain the validity of the unsteady flow model. In the third study, simulations are used for the numerical optimization of a configuration with multiple control surfaces. Here, the flexibility of the model is exploited in order to achieve a reduction of induced drag. Compari- son with experimental results shows that the numerical method attains adequate accuracy within the inherent limits of the potential flow model.

Finally, rather extensive aeroelastic simulations are performed for the ASK 21

sailplane. Time-domain simulations of a pull-up maneuver and comparisons

with flight test data demonstrate that, considering modeling and computational

effort, excellent agreement is obtained. Furthermore, a flutter analysis is per-

formed for the same aircraft using identified frequency-domain loads. Results

are found to deviate only slightly from critical speed and frequency obtained

using an industry-standard aeroelastic analysis code. Nevertheless, erratic results

for control surface hinge moments indicate that the accuracy of the present

method would benefit from improved control surface modeling and coupled

boundary layer analysis.

(6)
(7)

Dissertation

This thesis is based on a brief introduction to the area of research and the following appended papers:

Paper A

D. Eller. An Efficient Method for Time-Domain Low-Speed Aerodynamics.

TRITA/AVE 2005:40, Department of Aeronautical and Vehicle Engineering, KTH, December 2005.

Paper B

D. Eller and M. Carlsson. An Efficient Aerodynamic Boundary Element Method for Aeroelastic Simulations and its Experimental Validation.

Aerospace Science & Technology, 7(7):532-539, November 2003.

Paper C

D. Eller and S. Heinze. An Approach to Induced Drag Reduction with Experi- mental Evaluation.

AIAA Journal of Aircraft, 42(6):1478-1485, November 2005.

Paper D

D. Eller. Efficient Flight Mechanics Simulation of Elastic Aircraft Configura- tions. Presented at the International Forum for Aeroelasticity and Structural Dynam- ics, Munich, June 2005.

Paper E

D. Eller and U. Ringertz. Aeroelastic Simulations of a Sailplane.

TRITA/AVE 2005:41, Department of Aeronautical and Vehicle Engineering,

KTH, December 2005.

(8)
(9)

Contents

Preface 1

Abstract 3

Dissertation 5

Introduction 9

Fluid-structure interaction and aeroelasticity 11

Aeroelasticity: Problem or opportunity? . . . . 11

Contributing factors . . . . 13

Modeling for numerical simulation 14 Frequency- and time-domain analysis . . . . 15

Structural models and flight mechanics . . . . 15

Time-domain unsteady aerodynamics . . . . 16

Frequency-domain aerodynamics . . . . 18

Current method in comparison . . . . 19

Summary of research results 20 Discussion 21 Model preparation . . . . 21

Scope of validity of the flow model . . . . 22

Computational cost . . . . 23

Future work 23

References 25

Division of work between authors 29

Appended papers

(10)
(11)

Introduction

On a flight from Orlando to Cedar Keys, the pilot of a Beech 35 began a descent from the cruise altitude of 2600 m when approaching the destination airfield. Like many of the over 15 000 Beech 35 built during a 35-year long production, the aircraft was of the original design with a V-shaped tail as shown in Figure 1. At this time, this particular plane had been in service for 45 years and flown 7669 hours. When descending through 1400 m altitude at 260 km/h,

Figure 1: An early production model of the Beech 35 (1947).

well below the maximum permitted flightspeed of the aircraft, the pilot noted what he thought was turbulence. He reduced speed, and continued the descent.

After landing, substantial damage was detected near the tail, where skin sheets had been pulled off the fuselage bulkheads. The investigation performed by the National Transportation Safety Board (NTSB) concluded that the damage was caused by control surface flutter of the V-tail. Even the probable cause of the tail flutter could be identified, as stated in the incident report [1]:

”The skin of both elevator assemblies was made of magnesium

and paint drips flowing towards the trailing edge were noted on

both elevators. The shop manual in effect at the time of paint-

ing indicates, ’when the elevator/rudder control surface is being

repainted, suspend it by the trailing edge so that excess paint will

drain towards the leading edge.’ [...] Both elevators were then

removed for static balance check using a triple beam scale which

determined that the left was found to have an underbalance of 21.47

inch pounds and the right was found to have an underbalance of

(12)

10 D. Eller

21.38 inch pounds. Specification for balance is 16.8 to 19.8 inch pounds.”

It may appear rather surprising that paint, incorrectly applied, can lead to substantial structural damage. Unfortunately, the above example is not excep- tional. For the time between 1996 and 2002, five additional flutter incidents were reported with this particular aircraft type only, 1 all involving more or less severe structural damage.

How can paint drips in the wrong place cause so much damage? Figure 2 shows a section through a wing to which a trailing edge flap is attached, an arrangement similar to the tail of the Beech 35. Consider that, perhaps caused by turbulence, the wing bends up a little, corresponding to some upward ac- celeration a. An inertial reaction force F acts on the mass of the flap, for

Figure 2: Illustration of flap mass coupling.

now assumed concentrated in the center of gravity of the flap. Since the cen- ter of gravity is offset by a distance s from the hinge line around which the flap rotates, the reaction force entails a hinge moment. For a positive s, the resulting moment will lead to an increase of the flap deflection δ, which in turn generates an increase in lift force. While the wing is already temporarily ac- celerating upwards because of the turbulence, an upward-pointing aerodynamic force is generated which further amplifies the upward acceleration of the wing.

The amount of this amplification depends, among other factors, on the flight speed, the distance s and the mass of the flap (the product is the underbalance mentioned above) and the size and stiffness of the corresponding wing. The air- craft manufacturer therefore imposes limits on the underbalance, which must be checked after any modification, and possibly corrected by attaching lead weights to the forward part of the flap. In the case of the Beech 35, the weight of the excess paint at the trailing edge of the flap caused the underbalance to increase beyond the safe maximum.

1

NTSB incidents ATL97LA027, CHI98LA125, CHI00LA238, FTW02LA129, SEA02LA116

(13)

Fluid-structure interaction and aeroelasticity

The above case is an example of what is collectively referred to as fluid-structure interaction problems, which generally involve a fluid exerting pressure forces on a flexible structure. As an effect, the structure deforms, so that the fluid flow is disturbed. Fluid-structure interaction encompasses a broad range of phenomena from the flow in elastic blood vessels to rocket engine vibrations. In the context of aeronautical engineering, the term ’aeroelasticity’ relates to a class of specific fluid-structure interaction phenomena described in the following section.

This thesis is mainly concerned with the modeling and computation of aero- dynamic forces required for the numerical simulation of aeroelastic interaction.

As often is the case for dynamic problems, experimental aeroelastic investigations can be difficult, expensive and in the case of flight testing, highly dangerous.

Numerical analyses can help to evaluate a design before a prototype or model is built, or to identify critical conditions which require testing. A sufficiently accurate analysis may even be useful for the purpose of design optimization.

Aeroelasticity: Problem or opportunity?

Unsteady aerodynamics is, on its own, still considered a rather challenging dis- cipline. Interaction with flexible structures does certainly not simplify matters.

Traditionally, aeroelasticity has therefore often been regarded as a difficulty which is only taken into account very late in the development process, when prototypes are available for testing. This procedure appears to still prevail among many smaller aircraft manufacturers, while, in some cases, efforts are undertaken to consider aeroelastic effects earlier and perhaps even exploit them for improved performance.

Certification regulations [2], even for small sailplanes [3], require that any certified aircraft is free of static divergence and flutter for the range of permitted flight speeds. These two classical aeroelastic problems are related to the stability of the coupled fluid-structure system.

Divergence Aerodynamic loads lead to deformation of the load-carrying struc- ture, which changes the external shape of the aircraft. If the deformed shape generates larger aerodynamic loads than those which caused the initial deforma- tion, the deformation will increase even more and eventually result in structural failure.

Flutter In contrast to divergence, which can be imagined to develop arbitrar-

ily slowly, flutter is a dynamic, oscillatory instability. In this case, the mass

distribution of the structure comes into play. A small change in aerodynamic

(14)

12 D. Eller

loads leads to an equally small change in deformation. Because of the inertia of the structure, the deformation does not occur instantly, but with a certain delay. In addition, the accelerated structure may overshoot the equilibrium of elastic and aerodynamic forces slightly. Depending on the relative magnitudes and delays of fluid loads, elastic and inertial forces, the resulting oscillation can be damped (stable) or amplified (unstable).

Control surface effectiveness So far, the discussion has been restricted to stability considerations, but some other aspects also deserve mentioning. The perhaps most relevant is control surface effectiveness. Trailing edge flaps, when deflected downward, generate an increase in lift, but also a nose-down twist moment. Due to the twist deformation of the wing, this moment results in a decrease of the local incidence, which in turn causes a local decrease in lift force. Therefore, the effectiveness of trailing edge control surfaces on flexible wings declines with airspeed. At some critical velocity (the reversal speed), the effect of the control surface even changes sign, which, at least for manually controlled aircraft, appears rather inconvenient.

Exploiting flexibility For control surfaces at the leading edge, however, the effect is the opposite – this can be exploited to achieve improved, instead of degraded, performance. In N ASA ’s Active Aeroelastic Wing program [4], the maneuverability of a fighter aircraft with highly flexible wing is improved sub- stantially by means of leading edge control surfaces. Efforts within the European Union project Active Aeroelastic Aircraft Structures aim at actively controlling wing deformations in order to reduce induced drag [5, 6, 7]. Since aeroelastic interactions often are rather difficult to grasp intuitively, the design of active aeroelastic structures routinely requires the use of numerical optimization [8, 9].

Aeroservoelasticity Large, modern aircraft are often fitted with a digital flight

control system (FCS). The flight control computer is fed with measured data,

such as the stick position, speed, acceleration and rotation rates, and deter-

mines the appropriate deflection of the control surfaces. Quite naturally, the

dynamics of the control loop need at least be included in the analysis of aeroe-

lastic stability. Provided that the control surface actuation supports sufficiently

fast deflection, the FCS can also be employed to improve dynamic aeroelastic

damping and even suppress flutter at velocities beyond the critical flight speed

[10, 11]. Moreover, the FCS can be used to reduce both structural loads and

passenger discomfort caused by gusts and turbulence [12, 13]. It is reasonable

to expect that many of the possible benefits of actively controlled aeroelastic

aircraft necessitate the use of integrated aeroservoelastic optimization [14].

(15)

All of the phenomena mentioned above can occur in attached, low-speed flow, so that the computational methods presented as part of this thesis could be applied in analysis or optimization. In transonic flow, or in the presence of large separations, numerous other aeroelastic problems can arise [15], which require much more complex methods for simulation [16].

Contributing factors

As mentioned before, the interaction of structural dynamics and fluid flow con- stitutes the most important aspect in aeroelasticity. For both disciplines, a variety of mathematical models exist, many of which are suitable for numer- ical solution. Models of the load-carrying structure can account for nonlin- ear behavior in terms of geometry and material properties, or assume linearity.

Computational methods for the flow may allow simulations involving nonlinear phenomena such as strong vortices or transonic shock-boundary layer interac- tion, or they may be restricted to linear, inviscid aerodynamics.

Before different levels of physical modeling in simulation methods are com- pared, some additional aspects should be considered, as demonstrated by the following documented flutter incidents.

• In May 2003, a Canadair CL-600 (see Figure 3) declared an emergency during climb after takeoff from Atlanta and landed safely after the pilots had realized uncommanded roll movement and severe vibration. Inves- tigation identified freeplay in linkages connecting hydraulic actuators to the ailerons leading to aileron flutter as cause of the incident (NTSB case ATL03IA090).

• Stender [17] describes in detail the case of a high-performance sailplane which cleared extensive flight flutter tests up to 300 km/h without prob- lems. Later, some individual, heavily used aircraft of the same type ex- perienced repeated wing flutter with strong participation of the inboard flaps at 160 km/h. Investigations showed that the critical flutter speed was determined by the amount of friction in the flap actuation mechanism.

• In December 1968, a Cessna 182 sustained considerable damage due to aileron flutter, caused by water which had frozen inside the control sur- face. The aileron broke off the wing, but the pilot managed to land unharmed (NTSB incident OAK69F0211).

The first example of the above demonstrates the relevance of freeplay in the

control system, which allows a control surface to move without restraining force

within a narrow angular band. Freeplay will most likely increase during the

(16)

14 D. Eller

Figure 3: Canadair CL-600 2B19.

operation of the aircraft, and therefore requires that analyses can account for a range of conditions [18].

Similarly, friction can decrease due to wear in bushings and linkages, which can lead to a critical reduction of mechanical damping. As shown by the second example, aeroelastic instabilities can hence even occur with thoroughly flutter- tested aircraft types. Finally, the last case exemplifies the importance of problem awareness regarding aeroelasticity. Most likely, the maintainer of the Cessna 182 would not have expected that a small amount of water penetrating the aileron could lead to loss of the control surface.

In all three cases, even the most advanced structural analysis or the most detailed flow model would not have helped to identify and resolve the problem in advance. For that, a robust analysis of the dynamics of the control system would have been necessary, including aeroelastic effects and reliable data on possible variations of friction and freeplay in service. Naturally, even a robust analysis benefits from a more detailed modeling of the physics involved.

Modeling for numerical simulation

Naturally, numerical simulations are only useful as far as they can accurately

reproduce the relevant physics involved in the real problem. Additionally, the

numerical analysis must be sufficiently efficient to be useful. The latter require-

ment makes simplifications in the mathematical model of the physical problem

inevitable. In the following, different modeling approaches for structural dy-

namics and unsteady aerodynamics are compared in terms of modeling detail

and computational cost. Finally, the computational method developed as part

of this thesis is characterized with respect to the same criteria.

(17)

Frequency- and time-domain analysis

To analyse multidisciplinary problems involving unsteady aerodynamics, struc- tural dynamics, flight mechanics and possibly control systems, two fundamen- tally different approaches can be taken. The perhaps more straightforward route is to perform a time-marching simulation were the coupled equations of motion for the particular problem are advanced by taking discrete steps in time. Starting with a well-defined initial condition, an artificial disturbance is introduced and the response of the simulated system is monitored. Such a procedure is very flexible, as almost any kind of dynamic behavior can be incorporated as long as the corresponding equations of motion are known. The drawback of time- domain simulations is that, for each operating point (e.g. dynamic pressure, Mach number), a new simulation must be performed. Therefore, a single time- domain simulation of a sufficiently long interval should not require excessive computational effort.

Alternatively, equations of motion which only contain linear time-invariant terms can be transformed to the frequency domain assuming (damped) har- monic motion. Linearity assures that solutions can be superimposed, which gives rise to a modal representation of the motion. The advantage of this ap- proach is that the equations of motion can be cast as an eigenvalue problem, the solution of which provides direct information on the stability of the system.

The frequency-domain approach excludes nonlinear physical effects, and the approximation of freeplay or friction forces, which have comparatively simple descriptions in the time-domain, becomes rather involved [19].

Structural models and flight mechanics

For frequency-domain aeroelastic analysis, structural models assuming linear elastic structural response are required. These are normally considered suffi- cient to determine the stability of the aeroelastic system. An advantage of linear structural models is that the deformation can be expressed as a superposition of oscillations described by eigenmode shapes, which entails a considerable re- duction of the number of degrees of freedom. Moreover, if the eigenmodes are determined for the unconstrained aircraft, structural deformations and rigid body motion are not coupled by inertial forces [20]. Thus, only aerodynamic loads provide coupling between structural and flight mechanics degrees of free- dom which simplifies the analysis.

The assumption of geometrically linear behavior is most likely often justi-

fied, although there are configurations such as some high-performance sailplanes

(Figure 4) and unmanned aerial vehicles (UAVs, see Figure 5) which may not

be accurately described by means of linear structural models. Aircraft such

(18)

16 D. Eller

Figure 4: High-performance sailplane eta landing. (www.eta-aircraft.de)

as these probably require structural finite element models accounting for the geometrically nonlinear response to loading. In some cases, the nonlinear struc- tural response may not only cause quantitative differences, but lead to entirely different behavior [21].

In the case of the NASA research aircraft Helios, shown in Figure 5, the wing bending deformations are so large that they affect flight mechanics. With strongly bent wings, the rotational moment of inertia with respect to pitch in- creases significantly, which causes an unstable coupled pitch-bending oscillation [22]. The fuel-cell equipped research aircraft shown in the photograph was lost when this instability was excited by moderate turbulence during a test flight in June 2003. Highly flexible configurations can obviously not be described by linear structural mechanics and fully un-coupled flight mechanics. In order to accurately represent the behavior of such aircraft, nonlinear models are required, hence necessitating time-domain simulations of the coupled system.

Time-domain unsteady aerodynamics

The most detailed computational method for unsteady flow currently in use is the Detached-Eddy Simulation (DES). With sufficiently small timesteps, DES is capable of resolving the features of strongly separated flow and has shown good agreement with wind-tunnel experiments and flight test data for cases with complex flow patterns at high angles of attack, [23, 24, 25]. Such impressive ca- pabilities come at a price: For the analysis of rigid configurations, computation times in the order of 10 000 processor-hours are reported for simulated physical time intervals of approximately 0.05 s [25, 26].

Flight conditions with mainly attached flow can be simulated by means

(19)

Figure 5: Solar-powered research aircraft Helios. (NASA photo ED03-0152-4)

of unsteady Reynolds-averaged Navier-Stokes (RANS) methods. This approach intentionally time-averages small-scale turbulence in order to provide efficient solutions for attached high-Reynolds number flow. Unsteady RANS simulations can therefore be performed with larger timesteps, leading to a reduction of computational effort by a factor of 5-10 compared to DES [25]. At this cost, DES and aeroelastic RANS calculations can only be performed for a few conditions of particular importance, and even that would require access to high-performace computing facilities.

Further reductions of the computational cost are possible when viscous terms are neglected, which leads to numerical solutions of the Euler and full potential equations. While both methods are still nonlinear, the simplifications of the physical flow model allow the use of much less detailed field meshes than necessary for viscous simulations. Furthermore, the reduced complexity of the governing equations leads to more affordable numerical solutions [27, 28]. In a recent article, Woodgate et al. state that a single-case aeroelastic transonic Euler solution for the Hawk fighter aircraft with a mesh of 610 000 points required 56 processor-hours, substantially less than the expected cost of RANS simulations [29]. However, the authors conclude that the resolution of the full aircraft mesh is insufficient and proceed by analysing a better resolved wings-only configura- tion.

Performing coupled simulations involving structural deformations is, how-

ever, not trivial [30]. Most of the difficulties are related to the proper defor-

mation of the aerodynamic field mesh as a function of the aircraft motion,

especially for large deformations [31, 32], and the energy-conserving transfer of

surface loads. This difficulty affects all of the methods mentioned above, which

are collectively referred to as Computational Fluid Dynamics (CFD).

(20)

18 D. Eller

Summarizing, the above field methods for time-domain aerodynamic anal- ysis provide solutions with varying trade-offs between fidelity and cost. Formu- lation in the time-domain requires multiple aeroelastic simulations at different flight conditions, each of which requires considerable computational effort. This and the substantial work associated with model preparation and mesh genera- tion [33] makes these methods less attractive for the purpose of design. How- ever, once the aerodynamic geometry is fixed and reliable structural models are available, advanced aerodynamic analyses can provide valuable data e.g. for the preparation of flight tests.

Frequency-domain aerodynamics

The most common linear frequency-domain aerodynamic methods in use are the Doublet-Lattice Method (DLM, [34]), which can be regarded as a frequency- domain version of Hedman’s vortex lattice method [35], and panel methods such as Z AERO [36] and A EREL [37], based on the same model of unsteady compressible small-disturbance potential flow [38]. In comparison to the DES or RANS methods, the effects of viscosity are neglected, and rotational flow is limited to an infinitely thin wake surface. In both variants, lifting surfaces are approximated by their mean planes and discretized using quadrilateral panels (as shown in Figure 6). Due to the much simpler discretization, coupling of struc-

X Y

Z

X Y

Z

Figure 6: DLM panel mesh for a sailplane.

tural deformation and aerodynamic mesh displacement is less involved than for the field methods described in the previous section, but still not trivial as soon as the topology of the structural model differs from the flat-plate approximation of the aerodynamic surface.

The computational effort for the DLM or similar frequency-domain meth-

ods is in general fairly small, provided that the geometry under investigation

(21)

can be discretized with a few hundred flat panels, which usually is the case.

Computation of the aerodynamic load matrices required for the solution of an aeroelastic stability problem, can typically be performed in about two processor- hours on a desktop computer.

For the design of control laws for flexible aircraft, a compact representation of the aeroelastic system is needed, preferably in time-domain state-space form.

Results in the Laplace domain must be transformed by means of approximations [39, 40]. Time-domain aeroelastic analyses, however, can be reduced to state- space form using system identification methods [41, 42].

Current method in comparison

The method implemented as part of the present thesis is based on the assump- tion of potential flow. Thickness effects are accounted for by means of a more realistic discretization of the three-dimensional body surface, for which an ex- ample is shown in Figure 7. Furthermore, a time-domain formulation is chosen

Figure 7: Surface mesh of a sailplane undergoing a 4g maneuver.

to allow for simulations including nonlinearities. The incompressible poten- tial flow model is well suited for general time-domain simulations involving nonlinear control system or structural response due to the comparatively low computational cost.

To further improve computational efficiency for high-resolution meshes of

complex geometries, a panel clustering approach [43] is employed in conjunc-

tion with iterative solution techniques [44]. In this way, the computational effort

for complex geometries with several ten thousand surface elements is quite mod-

erate, so that unsteady simulations can be performed on inexpensive computing

hardware. Moreover, the use of an unstructured triangular surface mesh enables

the application of fully automatic surface mesh generation, which is a part of

(22)

20 D. Eller

the present implementation. In this way, human modeling effort can be reduced significantly.

Summary of research results

The main part of the presented work concerns the development and refinement of the numerical method for the computation of unsteady aerodynamic loads.

Additionally, algorithms for simulations involving flight mechanics of elastic structures have been implemented. Paper A describes the mathematical models used and some of the more important implementation features. Furthermore, convergence properties and accuracy of the method are illustrated by small ex- ample cases.

The remaining four papers deal with applications of the current method to wind tunnel experiments and flight tests. Paper B presents early work on validation of the computation of unsteady aerodynamic loads for transient elas- tic deformations. Here, it has been shown that the implementation provides good approximations for the generalized aerodynamic forces for relatively fast motion. The computational effort involved in this type of simulation is found to be rather moderate.

In Paper C, combined experimental work and numerical simulations are performed in order to investigate if aeroelastic effects can be exploited by means of redundant control surfaces to achieve a reduction of induced drag. Numeri- cal optimization is applied to a simulation model in order to find an optimal configuration, which is then tested in the wind tunnel. As a result, it is demon- strated that relevant drag reductions are possible for aircraft that need to operate over a wide range of lift coefficients. Experiments show that induced drag can be reduced as predicted, but also highlight the importance of accurate computation of control surface loads.

While earlier work is primarily concerned with the simulation of flexible,

but constrained wind-tunnel models, Papers D and E deal with configurations

undergoing large-scale motion in addition to small elastic deformations. First,

a relatively simple model, free to roll in the wind tunnel, is investigated. The

model carries control surfaces both on the leading and trailing edge and has

also been used in a different study on optimal control by Carlsson and Cro-

nander [8]. As a result of the work in Paper D, it is shown that the current

simulation method provides accuracy comparable to Z AERO in aeroelastic time-

domain simulations. Again, the importance of accurate prediction of control

surface forces and moments is emphasized. Both the present implementation

and Z AERO were found to perform inadequately in this respect for the particular

low-speed wind-tunnel model.

(23)

Finally, a full aircraft configuration is considered in Paper E. A detailed finite element model of a two-seat pilot training sailplane has been developed and validated using extensive vibration test data. A pull-up maneuver involving substantial loads was performed with an instrumented test aircraft. Numeri- cal simulations of the maneuver using the recorded time history of the eleva- tor deflection as the only input showed good agreement with respect to the flight mechanics states. Furthermore, the possible use of the current numerical method for the frequency-domain solution of the aeroelastic stability problem is explored.

Discussion

The appended papers contain detailed discussions of the specific cases treated there. In addition, some more general aspects shall be mentioned in the follow- ing.

Model preparation

Due to the use of automatic mesh generation, an aerodynamic surface mesh for the implemented method can be generated with relatively little effort. Steady and unsteady aerodynamic data for a rigid configuration can hence be obtained fairly quickly. As an illustration, Figure 8 shows steady pressure results for a simple twin-engine configuration. Starting from a simple sketch, the model preparation process required about four hours until steady coefficients had been computed for two meshes with different resolution.

Figure 8: Surface plot of steady pressures for a generic light aircraft.

Rigid-body results may be useful on their own in certain applications, but

the present method’s primary contribution is the ability to treat flexible aircraft.

(24)

22 D. Eller

This requires a reasonably accurate structural model of the complete aircraft, which often is rather involved and time-consuming to create. Figure 9 shows, as an example, the complex network of beam-type reinforcements which are part of a finite element shell model of the sailplane considered in Paper E. Most of these

Figure 9: Local reinforcements included in a structural model.

local reinforcements were initially thought to be of little relevance to the global structural response. Model validation using vibration test data demonstrated, however, that an appreciable level of structural detail was essential for sufficient accuracy.

As soon as a credible finite element model is available, the effort for the preparation of an aeroelastic model for the present method is moderate. This can be a significant advantage if development resources are scarce and can contribute to the introduction of aeroelastic investigations early in the design process.

Scope of validity of the flow model

Although the limitations of the underlying flow model are well understood, it can be difficult to determine if the problem investigated can be accurately described using potential flow. In general, the method should give reasonable results for attached flow at low Mach numbers.

For light aircraft, UAVs or wind tunnel models, the Reynolds number may

be below 10 6 . In this flow regime, boundary layers can be thick in comparison

to the body dimensions, effectively modifying the contour of the body as seen

by the external potential flow. For accurate results, the potential flow solution

should be computed for the displaced body contour. This functionality is not

implemented yet.

(25)

Computational cost

In comparison to CFD, the computational cost associated with the current method is almost negligible. Steady solutions for full aircraft configurations can be obtained in less than one minute on a Linux PC. Unsteady aeroelastic solutions typically require about 30 seconds per physical timestep, where the size of the timestep is determined by the highest frequency of the structural oscillations to be resolved.

An important aspect of the computational efficiency of the current method is that it does not restrict the application of the simulation method to a small number of costly investigations. In contrast to more detailed, but expensive methods, it can realistically be used for aeroelastic optimization and control system design. These aspects allow the designer to take advantage of aeroelastic effects, rather than correct unexpected aeroelastic problems discovered late in the development process or even flight testing.

Future work

Further development in the near term could be directed at the simulation of aeroelastic configurations involving nonlinear control system dynamics. Such behavior is caused e.g. by stick-slip friction in the mechanical actuation, and its effect on aeroelastic stability is not well understood. To perform this type of simulation, the aerodynamic method must be improved to reliably provide accurate aerodynamic control surface hinge moments.

In order to improve accuracy at low Reynolds numbers, it may be beneficial to extend the analysis with a coupled boundary layer model. For light aircraft and UAVs in particular, the thickness of the boundary layer may reach substan- tial fractions of the wing chord, and can therefore alter the pressure distribution.

Accounting for the displacement effect of the boundary layer would probably

improve the prediction of aerodynamic moments and control surface efficiency.

(26)
(27)

References

[1] National Transportation Safety Board. Incident Report MIA99LA185, June 1999. Available at www.ntsb.gov.

[2] European Aviation Safety Agency. CS-25, Airworthiness codes for large aero- planes., October 2003. Available at www.easa.eu.int.

[3] European Aviation Safety Agency. CS-22, Certification specifications for sailplanes and powered sailplanes, November 2003. Available at www.easa.eu.int.

[4] E. Pendleton, D. Bessette, P. Field, G. Miller, and K. Griffin. The active aeroelastic wing flight research program: Technical program and model analytical development. Journal of Aircraft, 37(4):554–561, July 2000.

[5] S. Kuzmina, F. Ishmuratov, V. Kuzmin, and Yu. Sviridenko. Estimation of flying vehicle induced drag changing due to deformation of lifting surfaces.

In CEAS/AIAA/NvVL International Forum for Aeroelasticity and Structural Dy- namics, Amsterdam, June 2003.

[6] J. Simpson, J. Schweiger, and S. Kuzmina. Design of an adaptive wing shape control concept for minimum induced drag of a transport aircraft.

In CEAS/AIAA/NvVL International Forum for Aeroelasticity and Structural Dy- namics, Amsterdam, June 2003.

[7] D. Eller and S. Heinze. An approach to induced drag reduction with experimental evaluation. Journal of Aircraft, 42(6):1478–1485, December 2005.

[8] M. Carlsson and C. Cronander. Efficient roll control using distributed control surfaces and aeroelastic effects. Aeronautical Science and Technology, 9(2):143–150, March 2005.

[9] G. W. Reich, D. Raveh, and P. S. Zink. Application of active-aeroelastic- wing technology to a joined-wing sensorcraft. Jounal of Aircraft, 41(3):594–

602, May 2004.

[10] M. Andrighettoni and P. Mantegazza. Multi-input / multi-output adaptive active flutter suppression for a wing model. Journal of Aircraft, 35(3):462–

469, May 1998.

[11] D. Borglund and U. Nilsson. Robust wing flutter suppression considering frequency-domain aerodynamic uncertainty. Journal of Aircraft, 41(2):331–

334, March 2004.

(28)

26 D. Eller

[12] S. V. Thornton. Reduction of structural loads using maneuver load control on the advanced fighter technology integration (AFTI) mission adaptive wing. Technical Report TM-4526, NASA, 1993.

[13] S. Heinze and M. Karpel. Analysis and wind tunnel testing of a piezo- electric tab for aeroelastic control applications. In CEAS/AIAA/NvVL In- ternational Forum for Aeroelasticity and Structural Dynamics, Munich, June 2005.

[14] G. Bindolino, S. Ricci, and P. Mantegazza. Integrated servostructural opti- mization in the design of aerospace systems. Journal of Aircraft, 36(1):167–

175, January 1999.

[15] A. Cunningham. Buzz, buffet and LCO on military aircraft. In CEAS/AIAA/NvVL International Forum for Aeroelasticity and Structural Dy- namics, Amsterdam, June 2003.

[16] E. F. Sheta. Alleviation of tail buffeting of F/A-18 aircraft. Journal of Aircraft, 41(2):322–330, March 2004.

[17] W. Stender. Praxisnahe Absch¨atzungs- und Vorbeugungsm¨ oglichkeiten gegen die Flattergef¨ahrdung von Segelflugzeugen und kleinen Motor- flugzeugen. Technical Report IB 151-74/6, 151-74/20 and 151-74/22, DFVLR e.V., Braunschweig, October 1974. Part I-III.

[18] D. H. Baldelli, R. Lind, and M. Brenner. Robust aeroelastic match-point solutions using describing function method. Journal of Aircraft, 42(6):1597–

1605, November 2005.

[19] H. K. Khalil. Nonlinear Systems. Prentice Hall, 3rd edition, 2000.

[20] B. Etkin and L.D. Reid. Dynamics of Flight. John Wiley and Sons, 1996.

[21] J. A. Garcia. Numerical investigation of nonlinear aeroelastic effects on flexible high-aspect-ratio wings. Journal of Aircraft, 42(4):1025–1036, July 2005.

[22] T. E. Noll et al. Investigation of the Helios mishap, Volume I. Technical report, NASA Headquarters, January 2004.

[23] J. R. Forsythe, K. D. Squires, K. E. Wurtzler, and P. R. Spalart. Detached- eddy simulations of the F-15E at high alpha. Journal of Aircraft, 41(2):193–

200, March 2004.

(29)

[24] D. M. Schuster and J. E. Byrd. Transonic unsteady aerodynamics of the F/A-18E at conditions promoting abrupt wing stall. In 41st Aerospace Sciences Meeting and Exhibit, Reno, January 2003. AIAA. Available at ntrs.nasa.gov.

[25] J. R. Forsythe and S. H. Woodson. Unsteady computations of abrupt wing stall using detached-eddy simulation. Journal of Aircraft, 42(3):606–616, May 2005.

[26] R. Sivaji. Aerodynamic analysis of the joined-wing configuration of a high- altitude, long-endurance aircraft. Master’s thesis, University of Cincinnati, May 2004.

[27] C. Farhat, P. Geuzaine, and G. Brown. Application of a three-field nonlin- ear fluid structure formulation to the prediction of the aeroelastic parame- ters of an F-16 fighter. Computers and Fluids, 32:3–29, January 2002.

[28] L. Fornasier, H. Rieger, U. Tremel, and E. van der Weide. Time-dependent aeroelastic simulation of rapid maneuvering aircraft. In 40th AIAA Aerospace Sciences Meeting and Exhibit, January 2002. AIAA paper 2002-0949.

[29] M. A. Woodgate, K. J. Badcock, A. M. Rampurawala, and B. E. Richards.

Aeroelastic calculations for the Hawk aircraft using the Euler equations.

Journal of Aircraft, 42(4):1005–1012, July 2005.

[30] C. Farhat, K. Lesoinne, and P. LeTallec. Load and motion transfer algo- rithms for fuid-structure interaction problems with non-matching discrete interfaces: Momentum and energy conservation, optimal discretization and application to aeroelasticity. Computer Methods in Applied Mechanics and En- gineering, 157:95–114, 1998.

[31] C. Degand and C. Farhat. A three-dimensional torsional spring analogy method for unstructured dynamic meshes. Computers and Structures, 80:305–

316, February 2002.

[32] A. K. Slone, K. Pericleous, C. Bailey, and M. Cross. Dynamic fluid struc- ture interaction using finite volume unstructured mesh procedures. Com- puters and Structures, 80:371–390, March 2002.

[33] T. J. Baker. Mesh generation: Art or science? Progress in Aerospace Sciences, 41:29–63, 2005.

[34] E. Albano and W. P. Rodden. A doublet-lattice method for calculating

the lift distribution on oscillating surfaces in subsonic flow. AIAA Journal,

2(7):279–285, July 1969.

(30)

28 D. Eller

[35] S. G. Hedman. Vortex lattice method for calculation of quasi steady state loadings on thin elastic wings. Technical Report 105, Aeronautical Research Institute of Sweden (FFA), 1965.

[36] P. C. Chen, H. W. Lee, and D. D. Liu. Unsteady subsonic aerodynamics for bodies and wings with external stores including wake effort. Journal of Aircraft, 30(5):618–628, September 1993.

[37] V. J. Stark. The AEREL flutter prediction system. In ICAS-90-1.2.3, 1990.

[38] J. D. Anderson. Modern Compressible Flow. McGraw-Hill International, 3rd edition, 2003.

[39] M. Karpel. Design of active flutter suppression and gust alleviation using state space aeroelastic modeling. Journal of Aircraft, 19(3):221–227, March 1982.

[40] G. Pasinetti and P. Mantegazza. Single finite state modeling of aerodynamic forces related to structural motion and gusts. AIAA Journal, 37(5):604–612, May 1999.

[41] W. A. Silva. Discrete-Time Linear and Nonlinear Aerodynamic Impulse Responses for Efficient CFD Analyses. PhD thesis, The College of William and Mary in Virginia, October 1997.

[42] D. E. Raveh. Identification of computational fluid dynamics based un- steady aerodynamic models for aeroelastic analysis. Journal of Aircraft, 41(3):620–628, May 2004.

[43] W. Hackbusch and Z. P. Nowak. On the fast matrix multiplication in the boundary element method by panel clustering. Numerische Mathematik, 54(4):463–491, 1989.

[44] Y. Saad. Iterative methods for sparse linear systems. PWS Publishing, 1996.

(31)

Division of work between authors

Paper B

The numerical investigations were performed by D. Eller, while the experiments were prepared and managed by M. Carlsson. Experimental data was analyzed jointly by M. Carlsson and D. Eller and the paper was written by D. Eller with assistance by M. Carlsson.

Paper C

S. Heinze designed and manufactured the wind tunnel model, conducted the experiments, and developed the structural finite element model. D. Eller per- formed the aerodynamic modeling and the optimization study. The paper was written by D. Eller with support from S. Heinze.

Paper E

U. Ringertz conducted the vibration and flight testing, assembled the experimen-

tal data and provided the modified pk-method flutter solver. D. Eller created

the structural shell model and performed the aeroelastic analyses. The paper

was written by D. Eller with support from U. Ringertz.

References

Related documents

This chapter describes the guidelines that are meant to increase the test process maturity and thereby the testability and stability in software development projects at the

To measure performance of episodic memory, I used the memory tasks in Phase 1 and 2 where participants had to recognize a list of 40 words (20 words that had already been

For frequency domain data it becomes very simple: It just corresponds to assigning dierent weights to dierent fre- quencies, which in turn is the same as using a fre- quency

[r]

This thesis is based on the need to investigate the potential of both link time optimization (LTO) as a vehicle for solving postponed build system decisions and proper

As we can see from figure 4.2.1 the transmission line is connected to the ground in order to provide the short circuit so the signal will travel trough the transmission line till

a) System A: After running the simulation in Eurostag and Dymola, the voltage magnitude for the NGEN and the NLOAD nodes (Fig. 1) have been chosen as output for validation. As can

The PSVD approach by applying the PQRD-BC algorithm for approximate singular value decomposition (SVD) of polynomial matrices is compared to the approach of performing a set