• Nie Znaleziono Wyników

Prediction of Aeroelastic Limit-Cycle Oscillations Based on Harmonic Forced-Motion Oscillations

N/A
N/A
Protected

Academic year: 2021

Share "Prediction of Aeroelastic Limit-Cycle Oscillations Based on Harmonic Forced-Motion Oscillations"

Copied!
38
0
0

Pełen tekst

(1)

Delft University of Technology

Prediction of Aeroelastic Limit-Cycle Oscillations Based on Harmonic Forced-Motion

Oscillations

van Rooij, Anouk; Nitzsche, J.; Dwight, Richard DOI

10.2514/1.J055852 Publication date 2017

Document Version

Accepted author manuscript Published in

AIAA Journal: devoted to aerospace research and development

Citation (APA)

van Rooij, A., Nitzsche, J., & Dwight, R. (2017). Prediction of Aeroelastic Limit-Cycle Oscillations Based on Harmonic Forced-Motion Oscillations. AIAA Journal: devoted to aerospace research and development. https://doi.org/10.2514/1.J055852

Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

(2)

Prediction of aeroelastic limit-cycle oscillations

based on harmonic forced-motion oscillations

A.C.L.M. van Rooij and J. Nitzsche

German Aerospace Center (DLR), Institute of Aeroelasticity Bunsenstrasse 10, 37073 Göttingen, Germany

Anouk.vanRooij@dlr.de, Jens.Nitzsche@dlr.de

R.P. Dwight

Delft University of Technology, Faculty of Aerospace Engineering Kluyverweg 1, 2629 HS Delft, The Netherlands

R.P.Dwight@tudelft.nl

Aeroelastic limit-cycle oscillations (LCO) due to aerodynamic non-linearities are usually investigated using coupled fluid-structure interaction simulations in the time domain. These simulations are computationally expensive, especially if a large number of LCO solutions must be computed to study the Hopf bifurcation behaviour in the immediate surrounding of the flutter point. To facilitate such bifurcation parameter studies an adaptation of the well-known p-k flutter analysis method is proposed in this paper. In this frequency domain method, the first harmonic of the motion-induced unsteady aerodynamic forces is no longer assumed to be solely determined by constant-coefficient frequency response functions. Instead, the non-linear dependence on the oscillation amplitudes and the phase angle between the input degrees of freedom is additionally taken into account. Therefore, the first harmonic Fourier components of the aerodynamic forces are sampled and interpolated in advance. The LCO solution is then found iteratively. The proposed amplitude-dependent p-k method (ADePK) is applied to a classic two-degree-of-freedom spring-mounted airfoil system where the non-linear aerodynamic forces are computed from Euler simulations. Fluid-structure

This is an Accepted Manuscript of an article published in: AIAA Journal

Available online: https://arc.aiaa.org/doi/10.2514/1.J055852

(3)

interaction simulations are performed for validation of the method. Both methods show good agreement. Furthermore, the ADePK method is shown to be a useful tool to rapidly study structural parameter variations.

(4)

Nomenclature

A = generalised aerodynamic force matrix,

b = wing span, m

c = chord length, m

D = damping matrix

Dα = pitch damping, kgm2/s/rad

Dh = plunge damping, kg/s

fL = function of the lift

fM = function of the moment

~

f = aerodynamic force vector ~ˆ

f = complex amplitude vector of the aerodynamic force ~˜

fi = interpolated complex amplitude vector of the aerodynamic force

h = plunge displacement, m

Iα = mass moment of inertia (about the elastic axis), kgm2

K = stiffness matrix

Kα = torsional spring stiffness, Nm/rad

Kh = plunge spring stiffness, N/m

k = reduced frequency

kf = reduced frequency at flutter

k∗

α = uncoupled reduced natural pitch frequency

k∗

h = uncoupled reduced natural plunge frequency

L = lift, N

ˆ

L = complex-valued lift, N |L| = magnitude of the lift, N

M = moment, Nm

ˆ

M = complex-valued moment, Nm

|M | = magnitude of the moment, Nm M = mass matrix

(5)

m = mass, kg

n = time step level

p = complex-valued eigenvalue

q∞ = freestream dynamic pressure, Pa

q∞,ref = reference freestream dynamic pressure, Pa

Sα = static moment (related to the elastic axis), kgm

t = time, s

U∞ = freestream velocity, m/s

~ˆx = eigenvector or motion vector

~x = time-dependent motion vector

α = pitch angle, rad

∆h = plunge amplitude, m

∆α = pitch amplitude, rad

∆x1 = amplitude of the first degree-of-freedom of the van der Pol oscillator ∆x2 = amplitude of the second degree-of-freedom of the van der Pol oscillator

δ = amplification rate

ǫ = damping coefficient

θhα = complex amplitude ratio

|θhα| = magnitude of the complex-valued amplitude ratio

|θhα|f = magnitude of the complex-valued amplitude ratio at flutter

θLα = complex-valued ratio of the first harmonic of the lift to the pitch amplitude, N/rad

θM α = complex-valued ratio of the first harmonic of the moment to the pitch ampliutde, Nm/rad

θL|θhα| = complex-valued ratio of the first harmonic of the lift to the amplitude ratio, N

θM|θhα| = complex-valued ratio of the first harmonic of the moment to the amplitude ratio, Nm

|θx1x2| = magnitude of the complex-valued amplitude ratio of the van der Pol oscillator

φα = phase angle of the pitch angle, rad

φh = phase angle of the plunge displacement, rad

(6)

(φhα)f = phase difference between pitch and plunge at flutter, rad

µ = bifurcation parameter

φLh = phase of the lift w.r.t. the plunging motion, rad

φM α = phase of the moment w.r.t. the pitching motion, rad

ω = angular frequency, rad/s

ωα = uncoupled natural pitch frequency, rad/s

ωh = uncoupled natural plunge frequency, rad/s

ωold = angular frequency of the previous iteration, rad/s

I. Introduction

Limit-cycle oscillations are periodic solutions of non-linear aeroelastic systems. In contrast to linear flutter, the amplitude of an LCO does not grow unbounded, but remains constant due to the presence of a non-linearity of the involved forces. Limit-cycle oscillations are caused by either structural or aerodynamic non-linearities or a combination of both. Although their effect might not be immediately detrimental, as in the case of linear flutter, they can lead to fatigue of the aircraft’s wings on the long term. In other words, if non-linear flutter is tolerated, because an LCO of small amplitude occurs at freestream velocities larger than the critical flutter speed, the flutter problem turns into a fatigue problem. Furthermore, the aeroelastic system might become unstable already at velocities lower than the critical flutter speed, i.e. limit-cycle oscillations might appear below the (linear) flutter boundary. Linear theory, which is used to predict classical flutter, fails to predict these (non-linear) oscillations.

Several researchers [1–4] have studied limit-cycle oscillations that occur on the F-16 aircraft if external stores are applied. The driving mechanism of these LCOs has not yet been understood in full detail, due the complicated non-linear behaviour. Aerodynamic sources of non-linearity are thought to be related to shock wave dynamics, flow separation and/or boundary layer transition. However, analysing the aerodynamic sources of non-linearity in detail is difficult, because of the high computational costs involved in coupled fluid-structure interaction (FSI) simulations in the time domain. Moreover, time domain simulations are not suited for detailed investigations into

(7)

the bifurcation behaviour of LCOs. Unstable LCOs, for example, which are repelling boundaries in the phase space, can usually not be found directly from time domain simulations. Furthermore, often multiple nested LCOs exist. In that case an unstable LCO is accompanied by a stable LCO (of larger amplitude). However, this is not necessarily the case, i.e. there might exist unstable LCOs without stable LCOs, as observed in classical mathematics, see e.g. Kuznetsov [5]. The transition from fixed-point stable solutions to stable or unstable LCO solutions is mathematically called the bifurcation behaviour. Following Dowell et al. [6], Figure 1 shows the two types of Hopf bifurcation that can occur for LCOs induced by aerodynamic non-linearities. When unstable limit-cycle oscillations exist below the linear flutter speed and the LCO amplitude decreases with increasing freestream velocity, a so-called subcritical bifurcation occurs (as indicated by the red line in Figure 1). When only stable LCOs occur above the linear flutter speed (i.e. without unstable LCOs below the flutter boundary) and the LCO amplitude increases with increasing freestream speed, a so-called supercritical bifurcation occurs (as depicted by the blue line in Figure 1).

Figure 1 Two types of LCO as described by Dowell et al. [6] (full line - stable LCO solutions, dashed line - unstable LCO solutions)

To ease the investigations into the bifurcation behaviour several alternatives to time domain methods are used. One of these methods is the aeroelastic harmonic balance (HB) method as first described by Greco et al. [7] and Thomas et al. [8]. In this aeroelastic HB method, a HB flow solver, in

(8)

which the governing flow equations are transferred into the frequency domain by Fourier transform, is used. The aeroelastic equations of motions are then solved iteratively in the frequency domain. The aerodynamic forces are obtained from the HB flow solver at each iteration. Greco et al. [7] used to the frequency-domain transonic small disturbance equations to determine the aerodynamic forces. A Computational Fluid Dynamics (CFD)-based HB method solving the Euler equations was used by Hall et al. [9]. Thomas et al. [8, 10–13] have applied the aeroelastic HB method, in conjunction with the HB flow solver of Hall et al., for studying limit-cycle oscillations induced by aerodynamic non-linearities. Further improvements towards a fully coupled solution method have been made by Ekici and Hall [14] and Yao et al. [15]. Yao et al. [15] have shown that the results obtained with their HB method are in good agreement with those obtained from coupled time domain simulations. Although, these improvements result in significant computational time savings, the flow solver must be called during the iterations of the structural solver. Furthermore, the CFD solver must be adapted to solve the governing fluid dynamics equations in the frequency domain.

Recently, neural network-based machine learning methods have also been applied to predict limit-cycle oscillations [16–18] due to aerodynamic non-linearities. In these studies, reduced-order models (ROMs) have been built in the time domain which predict the aerodynamic forces as a function of the airfoil motion. Finally, these ROMs are used to solve the equations of motion in the time domain. Good agreement with time domain simulations [17, 18] or HB method solutions [16] is observed if the neural network is sufficiently trained.

Another approach is to extend the frequency domain-based p-k-method, well-known in the context of linear flutter analysis. Ueda et al. [19] first applied a modified version of the p-k method, in combination with the transonic small disturbance equations, for predicting LCOs. Next to the frequency-dependence of the aerodynamic forces, as in the classical p-k flutter method, in this extended p-k method of Ueda et al., the amplitude-dependence of the aerodynamic forces is taken into account. Ueda et al. did this using superposition of the amplitude- and frequency-dependent aerodynamic forces obtained from each of the two degrees of freedom (pitch and plunge for a typical airfoil section used in a classical flutter analysis). To compute the aerodynamic forces

(9)

for each degree of freedom, Ueda et al. made a quasi-steady flow assumption, which in only valid for low reduced frequencies (< 0.3). Satisfactory agreement with time domain simulations was obtained for stable small-amplitude LCOs (< 0.5◦). However, for larger LCO amplitudes, the

extended p-k method of Ueda et al. predicted unstable LCOs, whereas the time-marching method encountered numerical instability of the flow solver. He et al. [20] recently applied an extended version of the p-k method similar to that of Ueda et al. However, they used CFD to compute the amplitude- and frequency-dependent aerodynamic forces due to pitch or plunge (and hence dropped the low frequency assumption). The method of He et al. was successful if the non-linearity was weak. However, for stronger non-linearities, deviations compared to the reference time-domain solution (and the harmonic balance solution) were present. Finally, Somieski [21] applied an eigenvalue method for the computation of limit-cycle oscillations of an aircraft nose landing gear. This eigenvalue method is based on superposition of the amplitude-dependent forces as well. When multiple non-linearities are present, these are connected by linear dynamic relations. In other words, a certain amplitude relation is chosen, dependent on the frequency, to represent the amplitudes of the other non-linearities as a function of that of the first non-linearity. Comparison with time domain computations showed excellent agreement.

In contrast to the methods of Ueda et al., He et al. and Somieski, in this work an adapted p-k method is presented in which the amplitude dependency of the aerodynamic forces is taken into account by the use of harmonic forced-motion sampling, but without assuming superposability of the motion-induced aerodynamic forces. The LCO is assumed to oscillate with its first fundamental frequency only (as in [19–21]), i.e. the higher order harmonic components of the LCO and the aerodynamic forces are neglected. This assumption has been justified by the authors in a previous work [22]. The adapted p-k method, from now on called the amplitude-dependent p-k method or ADePK for short, presented here is similar to the HB method. However, in contrast to the HB method, in this work, the Fourier transform of the aerodynamic forces is applied only at the output of the CFD code (i.e. at the forces themselves). Hence, in that case, no adapted CFD solver is needed for application of the ADePK method. Furthermore, the CFD solver must not be called during the solution procedure of the aeroelastic equations of motion. Instead, the aerodynamic forces are

(10)

obtained from interpolation of the response surface generated from the results of harmonic forced-motion sampling. A further difference between ADePK and the HB method is the influence of the higher harmonics of the aerodynamic forces on the first harmonic of those forces. This influence is implicitly taken into account in ADePK, whereas in the HB method the first harmonic of the aerodynamic forces depends on the number of harmonics that is used. That is, in the HB method the convergence of the aerodynamic forces must be assessed in order to ensure that the first harmonic of those forces converges when the number of higher harmonics that is taken into account is increased. This must be ensured for each LCO amplitude and mode shape. Kholodar et al. [23] show that, depending on the structural model, there is a significant influence of the number of harmonics used on the LCO bifurcation behaviour.

The results of ADePK are compared to time domain simulations results, in order to validate the method. The test case used in this work is a two-DOF pitch/plunge airfoil system. The results of the presented amplitude-dependent p-k method will be compared to the methods of He et al. and Somieski (i.e. using superposition of the aerodynamic forces). Furthermore, a structural parameter variation is performed. That is, the structural frequency ratio is varied, similar to the investigations of Kholodar et al. [23].

Section II describes the test cases, the CFD code and the time and frequency domain methods. The results of ADePK for the two-DOF airfoil system are then shown and discussed in section III. Finally, conclusions are drawn.

II. Computational methods and set-up

In this work two computational methods have been applied, a time domain method, which is used a reference, and a frequency domain method, ADePK. Our frequency domain method is an extension of the well-known p-k method used for flutter computations. In the time domain, fluid-structure coupling has been applied for validation of the frequency domain method. This section first describes the CFD code and set-up, then the structural model and the equations of motion are presented. The FSI method used for validation is outlined in Section II C. After that, the conventional and amplitude-dependent p-k method are presented. Finally, the harmonic forced-motion oscillation simulations are addressed.

(11)

A. CFD code and set-up

To determine the aerodynamic lift and moment the DLR TAU-code [24] is used. This CFD code is a finite-volume, cell-vertex-based, unstructured, compressible solver for both the Reynolds-Averaged Navier-Stokes (RANS) and the Euler equations. For spatial discretisation a 2nd-order central scheme [25] is used. Temporal discretisation is realised by dual time stepping [26], where in order to integrate in physical time, the implicit 2nd-order accurate Backward Differencing Formula (BDF2) integration scheme has been used. At each physical time step, the governing equations are integrated explicitly by adding a so-called pseudo time derivative.

The airfoil used in this study is the supercritical NLR7301 airfoil with a blunt trailing edge [27]. Its design Mach number is 0.72 and its design lift coefficient is 0.60. The chord length of the airfoil is 0.3 m. This study only considers Euler simulations. An unstructured O-type mesh with 1135 points has been used for all CFD simulations shown in this work. However, the mesh resolution and grid convergence are not of interest for the validation of ADePK, since the same mesh has been used for both time and frequency domain calculations. The farfield boundary has been placed 100 chord lengths away from the airfoil. The time step size used for all unsteady simulations is 1 · 10−4s. This

corresponds to 769 time steps per oscillation period for a reduced frequency of 0.1 and to 128 time steps per oscillation period for a reduced frequency of 0.6. This time step size was found to give time step size independent results. The test case considered in this work is at a Mach number of 0.74 and a mean angle of attack of −1.5◦.

B. Equations of motion and structural model

The aeroelastic system considered in this work is a spring-mounted airfoil with two degrees of freedom; pitch and plunge. The elastic axis of the system is located at the quarter-chord point, where the aerodynamic forces are applied. The equations of motion of the system (as shown in [22]) are given by:

    m Sα Sα Iα     | {z } M     ¨ h ¨ α     | {z } ~¨ x +     Dh 0 0 Dα     | {z } D     ˙h ˙α     | {z } ~˙x +     Kh 0 0 Kα     | {z } K     h α     | {z } ~ x =     −L M     | {z } ~ f , (1)

(12)

where m is the mass, Iα the mass moment of inertia, Sα the static moment, Kh the plunge

stiff-ness, Kα the torsional stiffness, Dh is the plunge damping and Dα is the torsional damping. The

aerodynamic force vector ~f consists of the aerodynamic lift L and the moment M . The plunge dis-placement is denoted by h and the pitch angle by α. The structural parameters of the wind-tunnel model Dietz et al. [28] are used in order to have a realistic structural model. They are given in Table 1. It should be noted here that no comparison to the wind tunnel experiments of Dietz et al. [28] is intended.

Structural parameter Value Units

Wing span b 1.0 m

Chord length c 0.3 m

Mass m 26.268 kg

Mass moment of inertia (about the elastic axis) Iα 0.079 kgm 2

Torsional spring stiffness Kα 6.646 · 10

3

Nm/rad

Plunge spring stiffness Kh 1.078 · 10

6

N/m

Static moment (related to EA) Sα 0.331 kgm

Pitch damping Dα 0.0687 kgm

2

/s/rad

Plunge damping Dh 45.764 kg/s

Table 1 Structural parameters for the 2 DOF NLR7301 airfoil system

C. Fluid-structure coupling

For reference, limit-cycle oscillations are also computed in the time domain. As described in [22] the coupling is partitioned. To achieve so-called “strong” coupling of the CFD and the structural solver, the forces and displacements are exchanged between these solvers multiple times during each physical time step. Figure 2 shows a schematic diagram of this “strong” coupling. In this work the coupling has been applied at each pseudo time step. In other words, at a certain time step, the forces of the previous physical or pseudo time step are used to compute the new structural displacements (steps 1 and 2). These are then fed back to the CFD code to compute the new aerodynamic forces at the current pseudo time step (steps 3 and 4), which then lead to new displacements (step 5 and 2). These are again fed back to the CFD code to compute the new aerodynamic forces (steps 3 and

(13)

4). This process is repeated for each time step until an equilibrium is established. Then the solver advances to the next time step. The numerical time integration of the fluid-structural problem (1) is also performed using a BDF2 integration scheme. In order to obtain the desired mean angle of attack, the airfoil is trimmed by subtracting the steady aerodynamic forces at this (mean) angle of attack.

In order to study the bifurcation behaviour, the freestream velocity has been varied. In the time domain simulations, this variation has been performed by varying the static temperature at constant Mach number.

Figure 2 Schematic of the fluid-structure coupling in the time domain

D. Conventional p-k method

The conventional p-k method for classical flutter computations was developed by Hassig [29]. In this method a solution to the equations of motion (1) of the form:

~x(t) = ~ˆxept, (2)

is assumed. Here ~ˆx indicates the complex-valued eigenvector and p is a complex-valued eigenvalue, defined by:

p = δ + iω, (3)

(14)

fre-quency k is computed from the angular frefre-quency via k = ωc/U∞. The aerodynamic force vector ~f

is given by:

~

f (t) =f e~ˆpt, (4)

where f is the complex amplitude vector of the aerodynamic force. Assuming linearity, this force~ˆ vector can be written as a Generalised Aerodynamic Force (GAF) matrix A times the eigenvector ~ˆx:

f = A~ˆx. (5)

Substituting the assumed solution, (2)-(5), into the equations of motion (1) yields:

p2M~ˆx + pD~ˆx + K~ˆx = A (k) ~ˆx. (6)

This eigenvalue problem must be solved iteratively, since the GAF matrix is a function of the reduced frequency, which is in turn part of the sought solution p. The GAF matrix consisted of derivatives of the aerodynamic forces and moments with respect to the degrees of freedom (or the chosen generalised coordinates). More details on the conventional p-k method can be found in Hassig [29].

E. Amplitude-dependent p-k method (ADePK)

The conventional p-k method cannot be used for the prediction of limit-cycle oscillations. At least not in its original form, since the aerodynamic forces are no longer a linear function of the displacements, i.e. A (k) does not exist in the non-linear case. In ADePK again a solution to the equations of motion (1) of the form (2) is assumed, similar to the conventional p-k method (with p defined by (3)). For the aerodynamic response of the system only the complex-valued first harmonic component is considered, since the higher harmonics are not of interest (as will be explained in section II F), i.e. (4). Substituting the assumed solution (2)-(4) into the equations of motion (1) yields:

(15)

p2M~ˆx + pD~ˆx + K~ˆx =f~ˆk, ~ˆx. (7)

In the conventional p-k method the right-hand side vector is written as a GAF matrix times the eigenvector (see (5)). In the non-linear case, this is no longer allowed. The aerodynamic force vector is now not only a non-linear function of the frequency, but also a non-linear function of the amplitudes of both input degrees of freedom and the phase angle between the degrees of freedom. Therefore, ~ˆx is now called the motion vector and (7) must be solved iteratively, for example using Newton’s method.

In order to uniquely determine ~ˆx one of the amplitudes (either pitch or plunge) is pre-set. The motion vector ~ˆx becomes:

~ˆx =     θhα· c 1    · ∆α. (8)

Then (7) is solved for two unknowns: the complex eigenvalue p and the complex amplitude ratio θhα = (∆h/c)/∆α · eiφhα = |θhα| · eiφhα. This is done for each pre-set amplitude (∆α or ∆h, here

∆α has been used). Since the force vector depends on the frequency and on the mode shape, the equations of motion need to be solved iteratively. Hence, the following problem must be solved:

                   p2M~ˆx + pD~ˆx + K~ˆx =f~ˆk, ~ˆx, ∆α = constant, ~ˆ fk, ~ˆx≈f~˜i  k, ~ˆx, (9)

wheref~˜irepresents the interpolated aerodynamic force vector. Similarly as for the conventional p-k

method the following steps are performed to obtain p and the complex amplitude ratio:

1. Fix ∆α

(16)

3. Compute the aerodynamic force vectorf~˜iat this frequency ω = kU∞/c and ~ˆx = [θhα· c, 1]T· ∆α

(by interpolation)

4. Solve the system of equations (7) for p and θhα

5. Take new ω = ℑ(p) and ~ˆx = [θhα· c, 1]T· ∆α

6. Iterate steps 3-5 until converged

7. Switch to next ∆α

This procedure is repeated for several pitch amplitudes, see also Figure 3 shows a flow chart of the solution procedure used in the amplitude-dependent p-k method for a certain freestream velocity. The amplitude ∆α at which the amplification rate, i.e. δ = ℜ(p), becomes zero, is the LCO amplitude. The other LCO properties, i.e. the plunge amplitude ∆h, the magnitude of the amplitude ratio |θhα| and the phase difference between pitch and plunge φhα, are determined from ~ˆx and the

reduced frequency is determined from the imaginary part of p. ADePK as described here can be used to determine the LCO amplitude when f = [ ˆ~ˆ L, ˆM ]T is known in advance as a function of the

frequency, pitch amplitude, plunge amplitude and the phase difference. Here, ˆL and ˆM describe the complex-valued responses of the aerodynamic lift and moment. The solution procedure used here to solve (7) is similar to that of Thomas et al. [8]. However, it should be noted that the pitch amplitude is not pre-set at the LCO amplitude (as Thomas et al. [8] did) in this work. Instead, ∆α is fixed at several chosen values and the complex eigenvalue p is computed at each of these amplitudes. From the relation between ∆α and p the LCO amplitude is then determined (i.e. when δ = 0). Furthermore, in the procedure outlined above and in Figure 3 the freestream speed is chosen and the aerodynamic forces at this chosen freestream speed are computed from the aerodynamic forces at the reference velocity, see also Section III A 2. Hence, U∞is not part of the solution vector

(17)

Figure 3 Flow chart of ADePK algorithm

Figure 4 schematically presents two possible curves for δ as obtained from a computation with ADePK (similar to Figure 2 of [21]). The pre-set amplitude ∆α is depicted on the horizontal axis. A positive δ indicates an amplified motion and a negative δ indicates a damped motion. The blue curve shows a positive δ at low amplitudes, then δ becomes negative with increasing amplitude. The LCO that occurs at δ = 0 is stable, i.e. it is an attractor. The red curve shows two intersections with the horizontal axis. At the first intersection, δ changes from negative to positive with increasing amplitude, i.e. an unstable LCO occurs, whereas at the second intersection with the horizontal axis, δ changes from positive to negative with increasing amplitude, i.e. a stable LCO occurs. The unstable LCO is a repeller as indicated by the arrows.

δ stable LCO stable LCO unstable LCO − + ∆α

(18)

F. Harmonic forced-motion sampling and construction of the response surface

In order to determine the aerodynamic lift ˆL and moment ˆM at each combination of amplitudes, frequency and phase difference, a so-called response surface is built using harmonic forced-motion CFD simulations. Interpolation on this response surface is then applied during the iterations of ADePK.

Equations (10) and (11) represent the time signal of the pitch angle α and the plunge displace-ment h. Note that the motion contains no higher harmonics.

h (t) = ∆h · sin (ωt + φhα), (10)

α (t) = ∆α · sin (ωt). (11)

The phase difference φhαis computed by subtracting the phase of the angle of attack φαfrom that

of the plunge displacement φh, i.e. φhα= φh− φα. That is, when φhαis positive plunge leads pitch.

The response of the aerodynamic forces to this harmonic motion is given by:

L (t) = |L| · sin (ωt + φLh) + h.h.t., (12)

M (t) = |M | · sin (ωt + φM α) + h.h.t., (13)

where h.h.t. stands for higher harmonic terms, |L| indicates the magnitude of the lift, |M | the magnitude of the moment, φLh the phase of the lift with respect to the plunging motion and φM α

the phase of the moment with respect to the pitching motion. In ADePK, only the complex-valued first harmonic of the motion and the aerodynamic response are taken into account. Although there are significant higher harmonics in the aerodynamic response (as observed from coupled FSI simulations), the work they perform on the airfoil is negligible, since the higher harmonics in the motion of the structure were observed to be very small compared to the first harmonic component (see [22]). Furthermore, every harmonic only performs work on itself. Therefore, if no higher harmonics exist in the motion they can be neglected in the aerodynamic forces. If the LCO has significant higher order components it is expected that ADePK will fail to predict the correct LCO solution. However, these LCOs were not observed from CFD simulations and experiments with aerodynamic non-linearities only [28, 30–32].

(19)

In contrast to the harmonic balance method as outlined by Thomas et al. [8], the effect the higher harmonic components of the aerodynamic forces on the first harmonic component of the aerodynamic forces is implicitly taken into account in ADePK. In the HB method multiple harmonics must be taken into account in order to include the influence of the higher harmonic components. Hall et al. [33] show that there is a significant effect of these higher harmonics on the first harmonic of the unsteady pressure distribution if the pitch amplitude is not small (∆α = 1◦ in their case).

This means that in the HB method, the convergence of the first harmonic of the aerodynamic forces must be checked for each unsteady flow computation. As noted in the introduction, Kholodar et al. [23] have investigated the influence of the number of harmonics used for computation of the aerodynamic forces on the bifurcation behaviour of the LCO solution.

The response surface is obtained from simulations where pitch and plunge are simultaneously applied, i.e. ~ˆ f =     ˆ L ˆ M    =     fL(∆α, |θhα|, k, φhα) fM(∆α, |θhα|, k, φhα)    . (14)

Instead of applying harmonic forced-motion sampling in a four-dimensional parameter space, the parameter space can also be reduced to two dimensions when superposition of the aerodynamic forces is applied even though the oscillation amplitude is not small. Although superposition is actually not allowed, the resulting aerodynamic force is then determined by the superposition of so-called describing functions (DFs). A DF, which is a well-known concept in control theory, expresses the (aerodynamic) response as a function of the frequency and the oscillation amplitude of a single DOF [34]. In that case, the pitching and plunging motions are applied to the airfoil in separate simulations for several frequencies and amplitudes. The resulting responses in terms of the aerodynamic forces are then added to describe the combined pitch/plunge motion, i.e.

~ˆ f =     ˆ L ˆ M    =     fL(∆α = 0, |θhα| = ∞, k, φhα= 0) + fL(∆α, |θhα| = 0, k, φhα= 0) fM(∆α = 0, |θhα| = ∞, k, φhα= 0) + fM(∆α, |θhα| = 0, k, φhα= 0)    . (15)

In this approach, the aerodynamic forces due to the combination of a pitching and a plunging motion are not taken into account. However, the samples at various phase differences are not necessary in this case, reducing the computational effort w.r.t. the non-linear case (i.e. sampling in

(20)

four-dimensional parameter space). As noted in the introduction, He et al. [20] and Somieski [21] used superposition of DFs to obtain the aerodynamic response.

The samples of the harmonic forced-motion oscillation simulations are uniformly spaced in all four dimensions (reduced frequency, amplitude ratio, phase difference, pitch amplitude), i.e. a tensor-product grid has been used. For interpolation on the response surface cubic spline interpo-lation has been applied.

The amplitude-dependent p-k method as described here with the equations of motion (7) can in principle be extended to more than two degrees of freedom. However, in that case the complex-valued aerodynamic forces and moments will be a function of more than four variables. The dimension of the complex-valued aerodynamic response increases as n + 1 + n(n − 1)/2 with the number of DOFs n. Therefore, the response surface dimensions and hence the amount of sample points will increase dramatically when the number of DOFs is increased. Depending on the number of structural parameter variations that need to be investigated, the use of ADePK for more than two degrees of freedom might be computationally too expensive. However, the construction method for the response surface used in this paper is the most simple possible. The number of samples points might be significantly reduced when adaptive sampling methods or a-posteriori methods are used (see e.g. Badcock et al. [35], who presented such methods for a Mach number and frequency-dependent response surface). This might lead to a reduction in the computational work, such that a response surface of the same order of sample points as that used in this paper can be used to investigate a more than two DoF system for example. Further investigations into more efficient response surface sampling techniques are necessary in order to optimise the construction of the response surface.

III. Results and Discussion

This section shows the results obtained from ADePK. The results are validated using time-domain FSI simulations. First, the response surface is shown and discussed. Then the LCO bifurcation behaviour of the two-degree-of-freedom airfoil system is shown and discussed. Finally, ADePK is applied to study the bifurcation behaviour of the 2 DOF airfoil system when the structural frequency

(21)

ratio is varied.

A. Two degree-of-freedom airfoil system

In order to apply ADePK to this test case, the aerodynamic forces were computed using inviscid CFD simulations. Harmonic forced-motion simulations were performed that sample the parameter space spanned by reduced frequency, amplitude ratio, phase difference between pitch and plunge and pitch amplitude (see Section II F). The results are used to create a response surface, which is used for the determination of the aerodynamic forces in ADePK. For validation, the LCO amplitude has been determined from FSI simulations at several freestream velocities.

1. Response surface construction

To determine the range of the mode shape parameters and frequency required, flutter calculations were performed (using the conventional p-k method) with varying structural parameters. The flutter mode shape was extracted for each structural parameter combination. Since the LCO mode can be expected to be similar to the flutter mode shape, the range in which to select response surface samples is determined from the flutter mode shapes. For this study, the structural parameters as depicted in Table 1 were taken as a starting point. Table 2 shows the variations applied to the structural parameters. The structural damping was left unchanged. Figure 5 shows the resulting flutter mode shapes for all combinations of the structural parameters in Table 2. From Figure 5(a) and 5(b) it is seen that the phase difference at flutter (φhα)f varies between 0◦ and 180◦. However,

most of the samples have a phase difference smaller than about 160◦. The amplitude ratio at flutter |θhα|f varies from 0 to about 20, with the highest concentration at amplitude ratios below 5. The

(22)

Structural parameter Values Units

Mass m 10, 18, 26, 34, 42, 50 kg

Mass moment of inertia Iα 0.01, 2.008, 4.006, 6.004, 8.002, 10 kg/m 2

Torsional spring stiffness Kα 3.323 · 10 3 , 6.646 · 103 , 9.969 · 103 , 1.3292 · 104 , 1.6615 · 104 , 1.9938 · 104 Nm/rad Plunge spring stiffness Kh 5.39 · 10

5 , 1.078 · 106 , 1.617 · 106 , 2.156 · 106 , 2.695 · 106 , 3.234 · 106 N/m Static moment related to EA Sα 0.10, 0.68, 1.26, 1.84, 2.42, 3.0 kgm

Torsional damping constant Dα 0.0687 kgm

2

/s

Plunge damping constant Dh 45.764 kg/s

Table 2 Values of the structural parameters used for determination of the range of response surface samples 0 5 10 15 20 0 30 60 90 120 150 180 |θ| f ( φhα )f ( ° )

(a) Phase difference vs amplitude ra-tio 0 0.2 0.4 0.6 0 30 60 90 120 150 180 kf ( φhα )f ( ° )

(b) Phase difference vs reduced fre-quency 0 5 10 15 20 0 0.1 0.2 0.3 0.4 0.5 0.6 |θ| f kf

(c) Reduced frequency vs amplitude ratio

Figure 5 Flutter mode shape for various structural parameters

The sampling range of the reduced frequency k was therefore taken from 0 to 0.6. For |θhα|, values

between 0.1 and 4 were used for the response surface samples and for φhαvalues between 5◦and 150◦

were used. Pitch amplitudes from 0◦ till 5were selected. To determine the number of samples in

each direction of the response surface, a comparison of an interpolated response surface slice with a dense-sampled response surface slice obtained from forced-motion oscillation simulations was made. The number of samples in each direction was chosen based on sufficient agreement between the two slices. Table 3 displays the sample locations of the mode shape parameters for determination of the

(23)

response surface and Figure 6 shows an exemplary cut through the response surface in terms of the complex-valued lift and the complex-valued moment versus the mode shape parameters. In Figure 6 blue dashed lines have been used for cubic spline interpolation. The slices versus pitch amplitude and amplitude ratio have been normalised by ∆α and |θhα|, respectively. In order to normalise with

respect to |θhα|, the aerodynamic force (or moment) at |θhα| = 0 is first subtracted.

Mode shape parameter Values

Pitch amplitude ∆α (◦ ) 0, 0.1, 0.5, 1, 2, 3, 4, 5 Amplitude ratio |θhα| 0.1, 0.5, 1, 4 Reduced frequency k 0, 0.1, 0.2, 0.25, 0.3, 0.4, 0.5, 0.6 Phase difference φhα( ◦ ) 5, 10, 50, 100, 150

Table 3 Values of the mode shape parameters used for determination of the response surface of two-DOF NLR7301 airfoil system

(24)

0 2 4 59.46 59.47 59.48 59.49 59.5 59.51 |θhα| ℜ ( θL|θ h α | ) (N) (a) 0 50 100 150 350 400 450 500 φ(°) ℜ (L) ( N ) (b) 0 0.2 0.4 0.6 400 600 800 1000 k ℜ (L) (N) Time domain Spline interp. (c) 0 1 2 3 4 5 2.7 2.72 2.74 2.76 2.78x 10 4 ∆α (°) ℜ ( θLα ) (N/rad) (d) 0 2 4 128.4 128.5 128.6 128.7 |θhα| ℑ ( θL|θ h α | ) (N) (e) 0 50 100 150 −300 −250 −200 −150 φ hα(°) ℑ (L) ( N ) (f) 0 0.2 0.4 0.6 −300 −200 −100 0 100 k ℑ (L) (N) (g) 0 1 2 3 4 5 −7000 −6800 −6600 −6400 ∆α (°) ℑ ( θLα ) (N/rad) (h) 0 2 4 1.1 1.15 1.2 1.25 1.3 |θhα| ℜ ( θM|θ h α | ) (Nm) (i) 0 50 100 150 −4.5 −4 −3.5 −3 φ(°) ℜ (M) ( N m ) (j) 0 0.2 0.4 0.6 −6 −5 −4 −3 k ℜ (M) (Nm) (k) 0 1 2 3 4 5 −400 −300 −200 −100 ∆α (°) ℜ ( θMα ) (Nm/rad) (l) 0 2 4 −1.04 −1.03 −1.02 −1.01 −1 |θhα| ℑ ( θM|θ h α | ) (Nm) (m) 0 50 100 150 −9 −8.5 −8 −7.5 φ hα(°) ℑ (M) ( N m ) (n) 0 0.2 0.4 0.6 −20 −15 −10 −5 0 k ℑ (M) (Nm) (o) 0 1 2 3 4 5 −600 −500 −400 −300 ∆α (°) ℑ ( θMα ) (Nm/rad) (p)

Figure 6 Response surface at M = 0.74, ¯α = −1.5◦

, ∆α = 1◦

, φhα= 5 ◦

, |θhα| = 0.5, k = 0.3

Figure 6 shows that the real and imaginary parts of lift and moment varied most with reduced frequency as could be expected. The deviation with respect to the linear part of response surface is clearly seen in the direction of the pitch amplitude. The slice of the response surface for the amplitude ratio was found to be only slightly non-linear, since the variations in the normalised real and imaginary of the lift and moment are small (much smaller than in the amplitude direction). For the phase difference the response surface slice seems to have a sine-like shape.

(25)

the flutter case is not expected to be a problem in the amplitude ratio direction (0.1 till 4 instead of 0.1 till 20) for the structural frequency ratio considered in this work (see section III B). In total 1280 samples were used (see Table 3). From those 980 are the output of forced-motion oscillation simulations. The aerodynamic forces at the other 300 sample locations are either identically zero (at a pitch amplitude of zero) or they have been determined from quasi-steady values (at a reduced frequency of zero).

Figure 7 compares the magnitude and phase angle of the moment versus the phase difference in the non-linear case and in case of superposition of the DFs, for two different pitch amplitudes (∆α = 1◦ and ∆α = 5). The amplitude ratio |θ

hα| is 1 and the reduced frequency k is 0.3. For

both amplitudes the shape of the magnitude and phase of the moment is approximately sinusoidal. At ∆α = 1◦ the describing function shows small deviations from the non-linear case. At ∆α = 5

on the other hand, the describing function of the moment is not correctly predicted at all. For the lift the agreement at ∆α = 5◦ is better, but small deviations are present as well. The plunge only describing functions also have a larger relative change in the magnitude and phase angle of the aerodynamic moment than for the lift.

(26)

−50 0 50 100 150 200 7.5 8 8.5 9 9.5 10 10.5 φhα (°) |M| Superposition of DFs Non−linear

Spline interp. (non−lin.)

(a) |M |, ∆α = 1◦ −50 0 50 100 150 200 −125 −120 −115 −110 −105 −100 φhα (°) φ Mα ( ° ) (b) φM α, ∆α = 1 ◦ −50 0 50 100 150 200 30 35 40 45 50 φhα (°) |M| (c) |M |, ∆α = 5◦ −50 0 50 100 150 200 −160 −150 −140 −130 −120 φhα (°) φ Mα ( ° ) (d) φM α, ∆α = 5◦

Figure 7 Magnitude and phase angle of the moment versus the phase difference at two pitch amplitudes at M = 0.74, ¯α = −1.5◦

, |θhα| = 1, k = 0.3

Figures 7(c) and 7(d) suggest that applying superposition of the aerodynamic forces or moments might result in a significantly different aerodynamic response than the actual non-linear response. Therefore, the bifurcation behaviour predicted from a response surface that was obtained from superposition of the DFs should be treated with care, especially for larger amplitudes.

2. Bifurcation behaviour

Using the samples as specified in Table 3, ADePK has been applied to study the bifurcation be-haviour of LCO amplitude of the NLR7301 airfoil. The amplitude at which δ becomes zero has been determined for several freestream velocities. The aerodynamic forces at each freestream velocity have

(27)

been determined by pre-multiplication of the dynamic pressure q∞ and dividing by the reference

dynamic pressure at which the aerodynamic forces were computed q∞,ref, i.e. f = q~ˆ ∞/q∞,ref·f~ˆref.

This means that formally the results are “non-matched”, as no additional iterations are performed to match the reference velocity with the computed flutter and LCO solution bifurcation velocities. Figure 8 shows an example of the amplification rate versus the pitch amplitude at a freestream velocity of 197.42 m/s. At this U∞the LCO amplitude is 3.33◦.

0 1 2 3 4 5 −0.15 −0.1 −0.05 0 0.05 ∆α (°) δ

Figure 8 Amplification rate versus pitch amplitude at M = 0.74, ¯α = −1.5◦

, U∞= 197.42 m/s

Figure 9(a) shows the results from ADePK together with the results from fluid-structure interaction simulations. Since it takes a lot of computational effort to determine the LCO amplitude in the time domain, at each freestream velocity several FSI simulations have been performed, each sim-ulation with a different initial amplitude. From each of these simsim-ulations it has been determined whether the oscillations of the system were growing or decaying in amplitude. In this manner, the bounds between which the LCO amplitude should lie have been determined. The blue circles and the red squares depict the lower and upper bounds between which a stable LCO occurs, respectively. Furthermore, unstable LCOs have also been found from time domain simulations at various veloci-ties. These were found from FSI simulations that decay in amplitude for a certain initial amplitude and increase in amplitude for a higher initial amplitude. The blue diamonds and red pentagrams depict the lower and upper bounds of these unstable LCOs, respectively. The black triangles at

(28)

zero amplitude show the results of FSI simulations for which the amplitude decays towards zero (at all initial amplitudes tested). The solid blue and red lines in Figure 9 show the frequency do-main solution in the non-linear case and in case of superposition of DFs, respectively. The flutter velocity obtained from the p-k method (using cubic spline interpolation) is plotted in Figure 9(a) with a green diamond at a zero amplitude. Figure 9(c) shows the non-dimensional plunge LCO amplitude and Figures 9(d) till 9(f) show the other LCO mode shape parameters (amplitude ratio, reduced frequency and phase difference between pitch and plunge) as obtained from ADePK versus the freestream velocity. For comparison the FSI simulation results are also shown.

From the time domain simulations it is seen that the unstable LCO amplitude decreases with increasing freestream velocity for amplitudes up to approximately 1◦. For larger amplitudes the

stable LCO amplitude increases with increasing amplitude, i.e. the bifurcation is supercritical. Below a freestream velocity of approximately 196.78 m/s all FSI simulations decay towards zero amplitude, i.e. no LCOs occur. The agreement of the results of ADePK with the time domain results is good. Unstable LCOs are predicted at small amplitudes for both the non-linear and the superposition case. However, in case of superposition of the DFs the LCOs become stable already at about 0.3◦, whereas in the non-linear case the trend of the FSI results is followed and stable LCOs

occur only just below 1◦, see Figure 9(b). Then, with increasing amplitude, the LCOs become

stable, then unstable and then stable again. These nested LCOs were not observed from the FSI simulations. However, the deviations are small, the overall trend is correctly predicted and the scale at which the errors occur is very small (0.1 m/s). The variations in the bifurcation behaviour between the two different response surface set-ups are a result of differences in the response surface. The results for the non-dimensional plunge amplitude obtained from ADePK show good agreement with the reference time simulations, see Figure 9(c).

(29)

196.50 197 197.5 198 198.5 199 199.5 1 2 3 4 5 U (m/s) ∆α ( ° )

(a) Pitch amplitude

196.750 196.8 196.85 196.9 196.95 0.5 1 1.5 2 2.5 3 U (m/s) ∆α ( ° )

(b) Pitch amplitude (zoom)

196.50 197 197.5 198 198.5 199 199.5 0.01 0.02 0.03 0.04 0.05 0.06 0.07 U (m/s) ∆ h/c

(c) Non-dimensional plunge amplitude

196.5 197 197.5 198 198.5 199 199.5 0.845 0.85 0.855 0.86 0.865 0.87 0.875 U (m/s) | θ hα | (d) Amplitude ratio 196.5 197 197.5 198 198.5 199 199.5 3.5 4 4.5 5 5.5 U (m/s) φ hα ( ° )

(e) Phase difference

196.5 197 197.5 198 198.5 199 199.5 0.306 0.307 0.308 0.309 0.31 0.311 U (m/s) k (f) Reduced frequency 196.8196.85196.9196.95 2.6 2.8 U ∞ (m/s) ∆α ( ° )

Lower bound unst. LCO (FSI) Upper bound unst. LCO (FSI) Lower bound st. LCO (FSI) Upper bound st. LCO (FSI) No LCO (FSI)

Flutter point (spline interp., p−k) Spline interp. (ADePK)

Superpos. (spline interp., ADePK)

(g)

(30)

Figure 9(d) shows a supercritical bifurcation of the amplitude ratio with the freestream velocity. The amplitude ratio is slightly underpredicted by the ADePK method in comparison to the FSI results (for both the non-linear and the superposition case). This is caused by the fact that the amplitude ratio has been computed from the pitch and plunge amplitudes, hence small discrepancies in those two parameters will lead to more significant deviations in the ratio of these two parameters. However, it should be noted that the scale is small and the relative changes w.r.t. the amplitude ratio at flutter are small. The largest relative deviation of the amplitude ratio from ADePK is 0.25% to 0.56% for the non-linear case and 0.70% to 1.0% in case of superposition of DFs.

The phase difference decreases with increasing airspeed, meaning that the pitching and plunging motions tend to get more in phase. The small positive phase difference indicates that plunge leads pitch. The reduced frequency is monotonically decreasing with freestream velocity, because k is an inverse function of the velocity (i.e. k = ωc/U∞). Although it should be noted that the variation

of reduced frequency is minimal, because the velocity variation is small. Upon comparing the results from ADePK to the time domain results, it is concluded that the bifurcation behaviour is globally correctly predicted by ADePK. Small discrepancies can be attributed to interpolation errors. The non-linear results agree better to the time domain results than the results obtained with superposition of DFs. Especially for the amplitude ratio and the phase difference the relative deviations are large between the superposed DFs and the non-linear results. To obtain the combined pitch/plunge results 1280 samples were used (980 forced-motion simulations). The superposition results were obtained using the same samples as depicted in Table 3 except the phase difference samples, hence 256 samples were used. This means that a significant amount of computational time is saved when superposition of DFs is applied. Therefore, using a response surface obtained from superposition of DFs as suggested by [20, 21] is an alternative when computational resources are limited. However, when accuracy is important, ADePK with fully non-linear aerodynamic forces should be used.

The efficiency of ADePK in combination with a response surface (in order to compute the aerodynamic forces) has been addressed for a single case at a velocity of 197.42 m/s and a pitch amplitude of 1.29◦. (Note from Figure 8 that this is not the LCO amplitude.) The results are shown

(31)

in Figure 10. ADePK needs 12 iterations when used in combination with a response surface and 6 iterations when used directly, i.e. when the CFD solver is called during each iteration of ADePK (in order to compute the aerodynamic forces). The number of iterations shown on the horizontal axis of Figure 10 is the total number of iterations of the CFD solver necessary when the aerodynamic forces are computed directly. In case a response surface is used to compute the aerodynamic forces, the number of iterations on the horizontal axis is an average number of iterations. This average has been computed based on an LCO bifurcation diagram that is established at 12 freestream velocities, as for the time domain simulations, and for which 50 pitch amplitudes were used to determine the LCO amplitudes at each velocity. The resulting amount of iterations has been equally divided among the 12 iterations of ADePK. A total number of 795876 iterations is necessary when a response surface is used. When the aerodynamic forces are computed directly during the iterations of ADePK, a total number of 3708000 iterations is necessary. Hence, when (7) must be solved at 12 freestream velocities with 50 pitch amplitudes each, the use of a response surface is already beneficial. Furthermore, when the LCO amplitude is required at more velocities, as necessary for the detailed prediction of the LCO bifurcation behaviour, the construction of a response surface is more efficient than computing the aerodynamic forces from CFD directly. Although, it should be noted here that the results obtained when the CFD solver is called during the iterations of ADePK are implicitly matched in contrast to the results obtained from ADePK in combination with a response surface. However, for the test case considered in this work the fact that ADePK, in combination with a response surface, produces non-matched results does not influence the bifurcation behaviour since good agreement between the results from ADePK and the time domain results (which are implicitly matched) was obtained. In addition, when the CFD solver is called during the iterations of the equations-of-motion-solver, an HB-approach, as in e.g. [8], can be used, which would remove the necessity of solving (7) at 50 pitch amplitudes. This might be beneficial up to a certain number of freestream velocities and in case of fixed structural parameters. However, when structural parameter variations are of interest at some point ADePK is expected to outperform HB, especially in combination with a more efficient response surface. Furthermore, for the HB method, the convergence of the first harmonic of the aerodynamic forces must be checked for each LCO amplitude, as suggested in Section II F. Alternatively, a very

(32)

high number of harmonics can be used to guarantee the convergence of the first harmonic for all computations. As explained in Section II F, in ADePK, the first harmonic of the aerodynamic forces implicitly contains all influences of the higher harmonics and hence no convergence checks are necessary when ADePK is used. This is another benefit of ADePK in comparison to the HB method. However, a direct comparison of the two methods is necessary in order to quantify the potential gains in computational efficiency of ADePK.

0 0.5 1 1.5 2 2.5 3 3.5 4 x 106 10−8 10−7 10−6 10−5 10−4 10−3 10−2 10−1 100 Number of iterations | ω old − ω |

ADePK resp. surf ADePK direcly

Figure 10 Residual versus number of iterations at M = 0.74, ¯α = −1.5◦

, U∞ = 197.42 m/s,

∆α = 1.29◦

B. Structural parameter variation

To demonstrate the capabilities of the proposed amplitude-dependent p-k method, ADePK has been used to study a structural parameter variation. Due to the separation of aerodynamics and structure, the structural model parameters can be easily varied in ADePK. There is no need to set up a new response surface (that is only necessary when the aerodynamic conditions are changed). Hence, variations in the bifurcation behaviour of the LCO amplitude and in the type of bifurcation can be studied very fast once the response surface is available. Here the natural structural frequency ratio (SFR) has been varied, similar to Kholodar et al. [23, 36], who performed a SFR variation for the NACA64010A airfoil. The bifurcation of the LCO mode shape obtained when varying the natural structural frequency ratio ωh/ωα from 0.49387 to 1.20971, is shown in Figure 11. The

(33)

variation of the structural frequency ratio has been obtained by varying the plunge spring stiffness Kh. The response surface has been obtained from combined pitch/plunge motions.

0.45 0.65 0.85 1.05 1.25 180 200 220 240 260 280 300 320 ωhα U ∞ f (m/s)

(a) Flutter speed

1750 200 225 250 275 300 325 350 1 2 3 4 5 ∆α LCO ( ° ) U (m/s) (b) Pitch amplitude 1750 200 225 250 275 300 325 350 0.5 1 1.5 2 2.5 3 | θ hα | U (m/s) (c) Amplitude ratio 1750 200 225 250 275 300 325 350 50 100 150 φ hα ( ° ) U (m/s) (d) Phase difference 175 200 225 250 275 300 325 350 0.1 0.2 0.3 0.4 0.5 k U (m/s)

uncoupled reduced natural

plunge frequency k∗

h

uncoupled reduced natural

pitch frequency k∗

α

(e) Reduced frequency

0 1 2 3 0 50 100 150 φ hα ( ° ) |θhα| 1.2097 1.1547 1.0969 1.036 0.97114 0.90168 0.82641 0.74355 0.65021 0.54101 0.49387

(f) Phase difference vs. amplitude ratio

(34)

From Figure 11(a) it is observed that the flutter speed first decreases and then increases with in-creasing structural frequency ratio. Furthermore, Figure 11(b) shows that the bifurcation behaviour becomes subcritical when the structural frequency ratio increases. The limit-cycle oscillations are plunge dominated, with a large |θhα| (>> 1), when the frequency ratio is small and pitch dominated,

with a very small |θhα| (<< 1), for the larger SFRs. This was also observed by Kholodar et al.

[23]. The phase difference is close to zero for small frequency ratios and becomes very large for large SFRs. The reduced frequency increases with increasing structural frequency ratio. However, for ωh/ωα> 0.90168 the reduced frequency decreases. Upon comparing the frequency at flutter to the

uncoupled reduced natural plunge frequency k∗

h, it is seen that for the lowest SFRs, at which the

bifurcation behaviour is supercritical, the flutter frequency is larger than k∗

h, whereas for the other

SFRs, with a subcritical bifurcation behaviour, the flutter frequency is smaller than k∗ h.

This structural parameter variation has been performed using the 1280 samples, i.e. 980 forced-motion oscillation simulations have been performed, after which a response surface has been built. Hence, the computational effort to built a response surface suitable for predicting limit-cycle oscil-lations using ADePK is large. However, the use of time domain simuoscil-lations would have resulted in a much higher computational effort, especially since new simulations are required as soon as a structural parameter is changed. Therefore, ADePK is a suitable tool to systematically investigate the bifurcation behaviour of limit-cycle oscillations, especially when structural parameter variations are of interest.

IV. Conclusions

In this work an alternative method for the prediction of limit-cycle oscillations has been pre-sented. This amplitude-dependent p-k method takes into account the complex-valued first harmonic of the aerodynamic forces as a function of the frequency, amplitude, amplitude ratio and phase dif-ference between the degrees of freedom. The method developed has been verified and validated for a two-degree-of-freedom airfoil system.

For the two degree-of-freedom airfoil system a response surface for the aerodynamic forces has been built using harmonic forced-motion simulations. This response surface has been used to determine the aerodynamic forces during the iterations of the p-k method. From both the

(35)

time domain and frequency domain simulations a supercritical bifurcation is observed for large amplitudes. Close to the flutter speed unstable LCOs were predicted by both methods. Overall, the agreement between the LCO amplitude and mode shape obtained from both methods is good. Hence, ADePK can be used to predict limit-cycle oscillations. Furthermore, taking into account only the first harmonic component of the aerodynamic forces is sufficient for the LCOs observed in this work. Therefore, once a response surface has been built for a certain Mach number and mean angle of attack, structural parameter studies can be easily performed using ADePK, as demonstrated in this work. Furthermore, superposition of describing functions can save further computational time with respect to the fully non-linear amplitude-dependent p-k method - although some accuracy is lost. ADePK could be extended to more than two DOFs, although the dimension of the problem will increase with the number of DOFs n of the system according to n+1+n(n−1)/2. However, ADePK could be applied to systems with more than two DOFs in the following manner: first a classical flutter analysis is performed. From this analysis the two degrees of freedom that couple during flutter are identified. These two degrees of freedom are then used to predict limit-cycle oscillations with ADePK. Finally, in order to improve the efficiency of ADePK further investigations into the response surface construction method are necessary.

References

[1] Denegri, C., “Limit Cycle Oscillation Flight Test Results of a Fighter with External Stores,” Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 761–769. doi:10.2514/2.2696.

[2] Bunton, R. and Denegri, C., “Limit Cycle Oscillation Characteristics of Fighter Aircraft,” Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 916–918. doi:10.2514/2.2690.

[3] Dreyer, C. and Shoch, D., “F-16 Flutter Testing at Eglin Air Force Base,” 3rd Flight Testing Conference, No. AIAA-86-9819, Las Vegas, NV, USA, 1986, pp. 1–6. doi:10.2514/6.1986-9819.

[4] Chen, P., Sarhaddi, D., and Liu, D., “Limit-Cycle-Oscillation Studies of a Fighter with External Stores,” 39th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, No. AIAA Paper 98-1727, Long Beach, CA, USA, 1998, pp. 258–266. doi:10.2514/6.1998-1727. [5] Kuznetsov, Y., Elements of Applied Bifurcation Theory, Vol. 112, Springer-Verlag, 2nd ed., 1998, pp.

(36)

[6] Dowell, E., Edwards, J., and Strganac, T., “Nonlinear Aeroelasticity,” Journal of Aircraft, Vol. 40, No. 5, 2003, pp. 857–874. doi:10.2514/2.6876.

[7] Greco, P., Lan, C., and Lim, T., “Frequency domain unsteady transonic aerodynamics for flutter and limit cycle oscillation prediction,” 35th AIAA Aerospace Sciences Meeting & Exhibit, No. 97-0835, Reno, NV, USA, January 1997, pp. 1–11.

[8] Thomas, J., Dowell, E., and Hall, K., “Nonlinear Inviscid Aerodynamic Effects on Transonic Diver-gence, Flutter and Limit-Cycle Oscillations,” AIAA Journal , Vol. 40, No. 4, 2002, pp. 638–646. doi: 10.2514/2.1720.

[9] Hall, K., Thomas, J., and Clark, W., “Computation of Unsteady Nonlinear Flows in Cascades Using a Harmonic Balance Technique,” AIAA Journal , Vol. 40, No. 5, 2002, pp. 879–886. doi:10.2514/2.1754. [10] Thomas, J., Dowell, E., and Hall, K., “Modeling Viscous Transonic Limit Cycle Oscillation Behavior

Using a Harmonic Balance Approach,” 43rd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, No. AIAA 2002-1414, Duke University, Durham, NC, Denver, CO, USA, April 2002, pp. 1–9. doi:10.2514/6.2002-1414.

[11] Thomas, J., Dowell, E., and Hall, K., “Modeling Limit Cycle Oscillations for an NLR 7301 Airfoil Aeroe-lastic Configuration Including Correlation with Experiment,” 44th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, No. AIAA 2003-1429, Duke University, Durham, NC, Norfolk, VA, April 2003, pp. 1–9. doi:10.2514/6.2003-1429.

[12] Thomas, J., Dowell, E., and Hall, K., “Modeling Limit Cycle Oscillation Behaviour of the F-16 Fighter Using a Harmonic Balance Approach,” 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, No. 2004-1696, Palm Springs, CA, USA, April 2004. [13] Thomas, J., Dowell, E., and Hall, K., “Further Investigation of Modeling Limit Cy-cle Oscillation Behaviour of the F-16 Fighter Using a Harmonic Balance Approach,” 46th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Ex-hibit, No. 2005-1917, 2005.

[14] Ekici, K. and Hall, K., “Harmonic Balance Analysis of Limit Cycle Oscillations in Turbomachinery,” AIAA Journal , Vol. 49, No. 7, 2011, pp. 1478–1487. doi:10.2514/1.J050858.

[15] Yao, W. and Marques, S., “Prediction of Transonic Limit-Cycle Oscillations Using an Aeroelas-tic Harmonic Balance Method,” AIAA Journal , Vol. 53, No. 7, July 2015, pp. 2040–2051. doi: 10.2514/1.J053565.

[16] Balajewicz, M. and Dowell, E., “Reduced-order modeling of flutter and limit-cycle oscillations using the sparse volterra series,” Journal of Aircraft, Vol. 49, No. 6, 2012, pp. 1803–1812. doi:10.2514/1.C031637.

(37)

[17] Zhang, W., Wang, B., Ye, Z., and Quan, J., “Efficient Method for Limit Cycle Flutter Analysis by Nonlinear Aerodynamic Reduced-Order Models,” AIAA Journal , Vol. 50, No. 5, 2012, pp. 1019–1028. doi:10.2514/1.J050581.

[18] Mannarino, A. and Mantegazza, P., “Nonlinear aeroelastic reduced order modeling by recur-rent neutral networks,” Journal of Fluids and Structures, Vol. 48, 2014, pp. 103–121. doi: 10.1016/j.jfluidstructs.2014.02.016.

[19] Ueda, T. and Dowell, E., “Flutter Analysis Using Nonlinear Aerodynamic Forces,” Journal of Aircraft, Vol. 21, No. 2, February 1984, pp. 101–109. doi:10.2514/3.48232.

[20] He, S., Yang, Z., and Gu, Y., “Transonic Limit Cycle Oscillation Analysis Using Aerodynamic Describing Functions and Superposition Principle,” AIAA Journal , Vol. 52, No. 7, July 2014, pp. 1393–1403. doi: 10.2514/1.J052559.

[21] Somieski, G., “An Eigenvalue Method for Calculation of Stability and Limit Cycles in Nonlinear Sys-tems,” Nonlinear Dynamics, Vol. 26, 2001, pp. 3–22. doi:10.1023/A:1017384211491.

[22] van Rooij, A., Nitzsche, J., and Dwight, R., “Energy budget analysis of aeroelastic limit-cycle oscillations,” Journal of Fluids and Structures, Vol. 69, 2017, pp. 174–186. doi: 10.1016/j.jfluidstructs.2016.11.016.

[23] Kholodar, D., Dowell, E., Thomas, J., and Hall, K., “Limit-Cycle Oscillations of a Typical Airfoil in Transonic Flow,” Journal of Aircraft, Vol. 41, No. 5, 2004, pp. 1067–1072. doi:10.2514/1.618.

[24] Schwamborn, D., Gerhold, T., and Heinrich, R., “The DLR TAU-code: Recent applications in re-search and industry,” European Conference on Computational Fluid Dynamics (ECCOMAS CFD), The Netherlands, September 2006.

[25] Jameson, A., Schmidt, W., and Turkel, E., “Numerical Solutions of the Euler Equations by Finite Volume Methods Using Runge-Kutta Time-Stepping Schemes,” 14th AIAA Fluid and Plasma Dynamics Conference, AIAA Paper 1981-1259, June 1981.

[26] Jameson, A., “Time Dependent Calculations Using Multigrid, with Applications to Unsteady Flows Past Airfoils and Wings,” AIAA Paper 91-1596 , June 24-26 1991.

[27] Zwaaneveld, J., “NLR 7301 airfoil,” EXPERIMENTAL DATA BASE FOR COMPUTER PROGRAM ASSESSMENT - Report of the Fluid Dynamics Panel Working Group 04 , No. AGARD-AR-138, 1979, pp. A4–1–A4–22.

[28] Dietz, G., Schewe, G., and Mai, H., “Experiments on heave/pitch limit-cycle oscillations of a super-critical airfoil close to the transonic dip,” Journal of Fluids and Structures, Vol. 19, 2004, pp. 1–16. doi:10.1016/j.jfluidstructs.2003.07.019.

(38)

[29] Hassig, H., “An Approximate True Damping Solution of the Flutter Equation by Determinant Iteration,” Journal of Aircraft, Vol. 8, No. 11, 1971, pp. 885–889.

[30] Schewe, G., Mai, H., and Dietz, G., “Nonlinear effects in transonic flutter with emphasis on mani-festations of limit cycle oscillations,” Journal of Fluids and Structures, Vol. 18, 2003, pp. 3–22. doi: 10.1016/S0889-9746(03)00085-9.

[31] Dietz, G., Schewe, G., and Mai, H., “Amplification and amplitude limitation of heave/pitch limit-cycle oscillations close to the transonic dip,” Journal of Fluids and Structures, Vol. 22, 2006, pp. 505–527. doi:10.1016/j.jfluidstructs.2006.01.004.

[32] van Rooij, A., Nitzsche, J., Bijl, H., and Tichy, L., “Limit-cycle oscillations of a supercritical airfoil,” International Forum on Structural Dynamics and Aeroelasticity (IFASD), Bristol, UK, June 2013. [33] Hall, K., Thomas, J., and Clark, W., “Computation of Unsteady Nonlinear Flows in Cascades Using a

Harmonic Balance Technique,” AIAA Journal , Vol. 40, No. 5, May 2002, pp. 879–886.

[34] Gelb, A. and Velde, W. V., Multiple-input describing functions and nonlinear system design, McGraw-Hill, 1968, Chapter 1: Nonlinear systems and describing functions.

[35] Badcock, K., Timme, S., Marques, S., Khodaparast, H., Prandina, M., Mottershead, J., Swift, A., Ronch, A. D., and Woodgate, M., “Transonic aeroelastic simulation for instability searches and uncer-tainty analysis,” Progress in Aerospace Sciences, Vol. 47, No. 5, 2011, pp. 392–423.

[36] Kholodar, D., Thomas, J., and Dowell, E., “A Parameter Study of Transonic Airfoil Flutter and Limit Cycle Oscillation Behaviour,” AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, No. 2002-1211, Denver, CO, USA, April 2002, pp. 1–16.

Cytaty

Powiązane dokumenty

W kolejnych rozdzia- łach znajdziemy zarówno informacje dotyczące zakonu kanoników regularnych late- rańskich, dziejów ich krakowskiego klasztoru, roli książki w życiu zakonu,

Sięgając po współ- czesną prozę, odnoszę jednak w niektórych wypadkach wrażenie, że przy okazji stało się też coś wręcz odwrotnego: nie tyle autorzy czy autorki

Na niebie ukazuje się straszliwy Perkunas i uderzeniem pioruna zapala stos na znak,że ofiarę przyjmuje.Stos z żubrem szybko płonie, a wbita w ziemię maczuga Ryngolda

Na apel odpowiedzieli tym razem Louis Jolicoeur (Université Laval, Québec) i Magdalena Nowotna (INALCO, Paryż), uznani specjaliści z zakresu przekładu literackiego, oraz

Strukturalne uwarunkowania opierają się raczej na kryteriach ewolucyjnego sukcesu i po- rażki, które pozwalają na podpartą binarnie au- toselektywność w procesach komunikacji, niż

In reviewing the three language-teacher-affiliated sociocultural theories, the psychological premises found in literature on the subject of individual differences,

zacji wszystkich szjtół 21. Za jeden z pierwszych etapów reformy szkół pT- jarskich członkowie delegacji uznali „przedsięwziąć negocjację w Rzymie, żeby zgromadzenie

In a stated choice experiment, the choices for three ISA implementation strategies (mandatory ISA, voluntary ISA with purchase subsidy and voluntary ISA with annual tax cuts)