12
American Institute of Aeronautics and Astronautics 1 STMA STUDY OF AIRFOIL NONLINEAR INTERACTION WITH UNSTEADY FLOW V. V. Golubev* and R. R. Mankbadi Embry-Riddle Aeronautical University Daytona Beach, FL 32114 ABSTRACT A computational method of Space-Time Mapping Analysis (STMA), developed in Ref. [1], is applied to nonlinear, inviscid computation of unsteady airfoil response to an upstream flow with a finite amplitude of unsteady perturbation. The unique feature of the method is that it solves the unsteady problem as a steady-state one, by treating the time coordinate identically to space directions. Time-accurate predictions are achieved by using a high-order discretization scheme. Computed results for both unsteady aerodynamic and aeroacoustic airfoil responses are compared against numerical solutions from Refs [2] and [3-4]. INTRODUCTION This paper re-examines a recently developed numerical approach, first presented in Ref. [1] and referred to as the Space-Time Mapping Analysis (STMA) method, wherein a two- or three-dimensional unsteady problem is treated as a steady-state one in the three- or four- dimensional space-time. The approach was developed with an idea to extend the high-accuracy discretization schemes developed in Computational Aeroacoustics (CAA) for spatial derivatives, to the time derivative, thus achieving highly time-accurate computations with absolute stability. Moreover, in terms of the distributed- memory parallel processing, the method appears to provide additional flexibility of clustering space-time grid in regions of rapid solution changes. The problem of unsteady gust-airfoil interaction, examined in the current paper, provides a convenient benchmark to investigate some of the beneficial features of the STMA method. For example, traditional CAA methods were developed to capture the transient ___________________________ * Assistant Professor, Senior Member of AIAA Professor, Associate Fellow of AIAA. behavior of an unsteady flow by combining a high- accuracy spatial differencing scheme with a time- marching method, thus treating any unsteady problem as a transient problem and maintaining high time accuracy throughout the calculation process. On the other hand, the gust-airfoil interaction problem is an example of a periodic problem where the exact transient solution starting from the initial conditions may not be of interest. Instead, the long-term periodic solution is usually the desired output, with the excessive accuracy of the transient calculations being redundant. Thus, in the STMA approach, an unsteady marching problem in two spatial dimensions is transformed into a steady- state iterative problem in three dimensions. In the process, highly convergent iterative schemes from classical CFD can be applied, with a potential for increased computational accuracy by using better time derivatives, reduced CPU time because of less grid points in space-time and more efficient iterative methods, and improved parallel performance through larger block volumes on each processor and reduced synchronization needs during the iterative process. The paper first provides a review of governing equations for the gust-airfoil interaction problem, followed by a detailed account of issues in numerical formulation and implementation of the STMA method. Presented results and discussion are split in two parts. First, unsteady response solutions computed for various frequencies of a small-amplitude impinging gust are compared and validated against results obtained for the same cases in Refs [2] and [3-4]. Next, the analysis focuses on a single frequency, but variable amplitude of the gust, to examine the nonlinear response features. GOVERNING EQUATIONS The current analysis of the gust-airfoil interaction problem is based on the numerical solution to nonlinear Euler equations, written in the Cartesian coordinates as 10th AIAA/CEAS Aeroacoustics Conference AIAA 2004-3003 Copyright © 2004 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

[American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

  • Upload
    reda

  • View
    217

  • Download
    0

Embed Size (px)

Citation preview

Page 1: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

1

STMA STUDY OF AIRFOIL NONLINEAR INTERACTION WITH UNSTEADY FLOW

V. V. Golubev* and R. R. Mankbadi†

Embry-Riddle Aeronautical University Daytona Beach, FL 32114

ABSTRACT

A computational method of Space-Time Mapping Analysis (STMA), developed in Ref. [1], is applied to nonlinear, inviscid computation of unsteady airfoil response to an upstream flow with a finite amplitude of unsteady perturbation. The unique feature of the method is that it solves the unsteady problem as a steady-state one, by treating the time coordinate identically to space directions. Time-accurate predictions are achieved by using a high-order discretization scheme. Computed results for both unsteady aerodynamic and aeroacoustic airfoil responses are compared against numerical solutions from Refs [2] and [3-4].

INTRODUCTION This paper re-examines a recently developed numerical approach, first presented in Ref. [1] and referred to as the Space-Time Mapping Analysis (STMA) method, wherein a two- or three-dimensional unsteady problem is treated as a steady-state one in the three- or four-dimensional space-time. The approach was developed with an idea to extend the high-accuracy discretization schemes developed in Computational Aeroacoustics (CAA) for spatial derivatives, to the time derivative, thus achieving highly time-accurate computations with absolute stability. Moreover, in terms of the distributed-memory parallel processing, the method appears to provide additional flexibility of clustering space-time grid in regions of rapid solution changes. The problem of unsteady gust-airfoil interaction, examined in the current paper, provides a convenient benchmark to investigate some of the beneficial features of the STMA method. For example, traditional CAA methods were developed to capture the transient ___________________________ * Assistant Professor, Senior Member of AIAA † Professor, Associate Fellow of AIAA.

behavior of an unsteady flow by combining a high-accuracy spatial differencing scheme with a time-marching method, thus treating any unsteady problem as a transient problem and maintaining high time accuracy throughout the calculation process. On the other hand, the gust-airfoil interaction problem is an example of a periodic problem where the exact transient solution starting from the initial conditions may not be of interest. Instead, the long-term periodic solution is usually the desired output, with the excessive accuracy of the transient calculations being redundant. Thus, in the STMA approach, an unsteady marching problem in two spatial dimensions is transformed into a steady-state iterative problem in three dimensions. In the process, highly convergent iterative schemes from classical CFD can be applied, with a potential for increased computational accuracy by using better time derivatives, reduced CPU time because of less grid points in space-time and more efficient iterative methods, and improved parallel performance through larger block volumes on each processor and reduced synchronization needs during the iterative process. The paper first provides a review of governing equations for the gust-airfoil interaction problem, followed by a detailed account of issues in numerical formulation and implementation of the STMA method. Presented results and discussion are split in two parts. First, unsteady response solutions computed for various frequencies of a small-amplitude impinging gust are compared and validated against results obtained for the same cases in Refs [2] and [3-4]. Next, the analysis focuses on a single frequency, but variable amplitude of the gust, to examine the nonlinear response features.

GOVERNING EQUATIONS The current analysis of the gust-airfoil interaction problem is based on the numerical solution to nonlinear Euler equations, written in the Cartesian coordinates as

10th AIAA/CEAS Aeroacoustics Conference AIAA 2004-3003

Copyright © 2004 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

Page 2: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

2

++

=

+

+=

=

=++

)(

)(

,0

2

2

pEvpv

uvv

F

pEuuv

puu

E

Evu

Q

FEQ yxt

ρρρ

ρρρ

ρρρ

(1)

with the unsteady pressure calculated from

+−−= )(

21)1( 22 vuEp ργ (2)

In this work, the gust response is investigated for the cambered, thick Joukowksi airfoil, which requires recasting the equations in the generalized curvilinear coordinates. In the traditional time marching approach, implemented, e.g., in Refs [3-4], the employed transformation takes the form,

( )( )yx

yx,,

ηηξξ

==

and the resulting Euler equations can be written as,

0=++++ ηξηξ ηξηξ FFEEQ yyxxt (3) However, in the STMA process, the governing equations (1) should be transformed into curvilinear coordinates that are all functions of space and time, as follows:

ξ = ξ x, y, t( )η =η x, y,t( )τ = τ x, y, t( )

(4)

In this development, it is important to note that τ is a function of x, y, and t; thus τ is not, in general, a time-like variable. Substituting into Eq. (1), the following equations are obtained in the strong conservation form:

0=

++

+

++

+

++

η

ξ

τ

ηηη

ξξξ

τττ

JF

JE

JQ

JF

JE

JQ

JF

JE

JQ

yxt

yxt

yxt

(5)

Note that τ -coordinate is defined in space-time in this approach (τ = τ (x,y,t)), rather than the usual definition of τ = τ (t). Thus, there is no time-like variable in the STMA approach, and a standard time-marching approach cannot be used. Instead, an iterative method can be used to solve Eq. (5). In application to the unsteady gust-airfoil interaction problem, the periodic nature of the flow may be used to minimize the computational time and effort required to solve the test cases. Particularly, the mesh is designed to cover one period of the vortical gust in the time direction, with a periodic boundary condition applied at the time inflow and time outflow boundaries. Thus, the computed solution is driven directly to the final long-time periodic solution of interest, rather than expending effort in accurately resolving the initial transient solution.

GUST-AIRFOIL INTERACTION MODEL

In application to the gust-airfoil interaction problem, it can be assumed that the unsteadiness of the incident flow is superimposed far upstream on the mean flow in the form of the vortical, convected perturbation velocity, represented in the space of perturbation frequencies and wave numbers in the integral form,

'( , ) Re{ ( )exp[ ( )]}kk

u x t A x i k x tω

ω= ⋅ −∑∫ rr

rr r r (6)

The incident perturbation field is further considered to be of finite amplitude relative to the mean flow, to the level that one cannot assume a linear superposition of responses from each individual harmonic in (6). It is clear that while the very notion of perturbation may become less obvious, the approach is valuable in that it allows to determine the limits of validity for linear approach universally accepted in studies on unsteady aerodynamics and aeroacoustics. Thus, similar to Ref. [4], the airfoil response is studied for the following distribution of the incident gust perturbation velocity,

Page 3: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

3

)cos(

)cos(

22

22

tyxMv

tyxMu

gust

gust

ωβαβα

εα

ωβαβα

εβ

−+

+=

−+

+−=

(7)

where ε is the gust intensity relative to the mean flow, α and β are the gust wave numbers, and ω is the imposed gust frequency. The mean flow is defined far upstream from the airfoil as:

)(10

1

γργ

ρ

=

===

p

vMu

(8)

where M is the Mach number, and γ=1.4. The two-dimensional gust is considered with

kMkk 2,2,2 === ωβα , where k is the reduced frequency of the gust non-dimensionalized by the airfoil half-chord and the upstream mean flow velocity.

NUMERICAL IMPLEMENTATION The H2 Advanced Concepts, Inc., Mapping Analysis Research Code 1 (MARC1) has been developed to solve the STMA equations [1]. This solver uses the Tam and Webb 7-point 4th order Dispersion Relation Preserving (DRP) scheme [5], with the 10th order artificial dissipation [6] added in all coordinate directions. To obtain the solution of Eq. (5), a time-like term may be added to the equation, and, e.g., the third-order Runge-Kutta scheme of Jameson et al [7] can be used in combination with implicit residual smoothing and local time stepping to accelerate convergence. However, in application to the cases presented, a single iterative step was used with no implicit residual smoothing. Figure 1 provides illustration of the 3D computational domain (x,y,t), along with contours of the computed conservative variable ρv from the solution vector Q in Eq. (1) obtained in a case study for the gust reduced frequency k=1.0. The additional time coordinate points in the z-direction. The translucent contour plot thus shows gust evolution both in space and time. The (x,y)-planes correspond to slices of the (x,y,t)

Figure 1: Computational (x,y,t) domain with contours of instantaneous ρv –component of gust. k=1.0, ε=0.2.

Figure 2: (x,y) slice of the (x,y,t) domain in Figure 1. domain, with the airfoil located in middle of the plane (in Figure 2). Note that the same (x,y)-plane was used as the entire computational domain in Refs [3-4], where the conventional time marching was applied. As mentioned in Ref. [4], the C-grid for the (x,y)-plane was generated as a single block around a thick, cambered Joukowski airfoil using the GridPro commercial software [8]. At the beginning of the iterative, pseudo-time marching convergence process, the flow was initialized

Page 4: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

4

throughout the physical domain with the gust perturbation superimposed on the mean flow. At the inflow boundary of each (x,y) plane, the standard acoustic radiation condition was applied to the outgoing perturbation, defined, e.g., for X-component of flow velocity as gustboundaryBC uuuu −−= . At the outflow boundary of each (x,y) plane, Tam and Webb’s [5] condition was applied. In the current work, computations for three selected gust frequencies were run on two meshes corresponding to Nx*Ny*Nt=(605x240x33) and (433*125*17). The current status of the code only allows for computations on a singe processor, which requires significant computer memory. A typical case run on a 2.8 GHz, 2GB SDRAM processor was taking up to a week to fully converge.

RESULTS AND DISCUSSION The computations discussed in this paper were conducted for gust reduced frequencies k=0.1, 1.0 and 2.0. The unsteady response was investigated for a loaded, 12%-thick, 2%-cambered Joukowski airfoil at a two-degree angle of attack. In all computations, Mach number was fixed at M=0.5. Below, results for low-amplitude gust response, calculated for all three frequencies, are examined first. These solutions are compared and validated against results obtained for similar cases in Refs [2] and [3-4]. The analysis then focuses on a single gust frequency of k=1.0, and examines the nonlinear response features for gust intensities ranging from ε=0.02 to ε=0.4.

Figure 3: Mean pressure on the airfoil suction and pressure sides: comparison of FLO36 solution with STMA prediction.

Figure 4: RMS pressure on airfoil surface, k=0.1, ε=0.02.

Figure 5: RMS pressure on airfoil surface, k=1.0, ε=0.02.

Figure 6: RMS pressure on airfoil surface, k=2.0, ε=0.02.

Page 5: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

5

Figure 7: Distribution of |p’2| at R=1, k=0.1, ε=0.02.

Figure 8: Distribution of |p’2| at R=4, k=0.1, ε=0.02.

Figure 9: Distribution of |p’2| at R=1, k=1.0, ε=0.02.

Figure 10: Distribution of |p’2| at R=4, k=1.0, ε=0.02.

Figure 11: Distribution of |p’2| at R=1, k=2.0, ε=0.02.

Figure 12: Distribution of |p’2| at R=4, k=2.0, ε=0.02.

Page 6: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

6

Low-Amplitude Gust Response for k=0.1, 1.0, 2.0 The STMA results obtained for the gust amplitude of ε=0.02 are shown in Figures 3-12. The mean pressure distribution on the airfoil surface is practically identical for all gust frequencies, and compares well with FLO36 predictions in Figure 3 (note that the results from the latter are also used as input in GUST3D code). In Figures 4-6, the predicted airfoil surface RMS unsteady pressure is compared against GUST3D results, as well as solutions computed in Refs [3-4]. Note that the latter match closely with results predicted by MARC1, whereas GUST3D results start to show deviation from the both time-accurate solutions at k=2.0. This could be related to the known loss of accuracy in GUST3D at higher reduced frequencies, usually more pronounced in the far field predictions as mentioned below.

The directivity plots for the non-dimensional acoustic intensities are shown in Figures 7-12, for each frequency calculated at distances R/c= 1 and 4 from the airfoil centerpoint. GUST3D results once again show rather significant deviations from both the current prediction and those obtained in Refs [3-4]. On the other hand, a similar deviation for this loaded airfoil response was observed before in Ref. [4] and references therein, and was attributed to GUST3D problems in the far field. To some extent, this problem may have been resolved in the improved GUST3D version which results will be reported in Ref. [3] proceedings; however, the relevant data has not yet been made available for the current work. Importantly, the obtained directivity predictions are in very good agreement with results from Refs [3-4]. One may note that at k=2.0, the far field directivity (R=4) starts to exhibit some distortion in the lobes. But this result was obtained on the grid 605x240x33, the finest used in current computations. Ref. [3] actually failed to achieve any reasonable far-field solution for this frequency on such a mesh due to poor gust resolution, and Ref. [4] result, shown for comparison, has been obtained only after employing a finer 515x971 mesh. The reader is referred to Ref. [4] for more discussion on the issue of grid resolution. High-Amplitude Gust Response for k=1.0 The airfoil unsteady response results obtained at various amplitudes of the incident gust for the reduced frequency k=1.0, are now reviewed. Note that most of these results were first reported in Ref. [9]. In

computations, the gust amplitude was ranging from ε=0.02 to ε=0.4. Note that along the airfoil surface, the aerodynamic response contains both the mean (due to the airfoil camber and angle of attack) and unsteady perturbation (due to the incident gust) pressure components. The results for the mean pressure distribution on the suction and pressure sides, computed for different gust amplitudes, are compared in Figure 13 with computations from FLO36 mean flow solver. Note that the STMA results are in good agreement with FLO36 solution, with some deviation observed near the leading edge for the highest gust intensity of ε=0.4. In fact, the STMA-produced pressure contours illustrated in Figures 14-15 for ε=0.02 and 0.4, respectively, show the presence of a shock and a small supersonic bubble in the case of the maximum gust amplitude. This produces characteristic wiggles observed in Figure 13 near the shock region. An FFT analysis of the pressure solution did not, however, indicate any energy transfer between the mean flow and the higher frequency harmonics in this case.

Figure 13: Airfoil mean pressure on the suction and pressure sides: comparison of FLO36 solution with STMA predictions at ε=0.02, 0.2 and 0.4. Figures 16-19 focus on the airfoil unsteady pressure RMS and FFT. General comparison of GUST3D RMS pressure predictions with STMA computations for different gust intensities is provided in Figure 16. The response amplitudes are scaled to the same input gust amplitude to allow comparison. In all cases, the agreement is excellent between STMA and GUST3D solutions, with small deviations observed near the trailing edge due to a local loss of numerical accuracy.

Page 7: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

7

Remarkably, the moderate deviations in the mean pressure did not propagate into much larger deviations in the unsteady pressure response. Moreover, the base FFT harmonic shown in Figure 17 appears closer to GUST3D prediction, which is particularly noticeable for the highest gust intensity. This may indicate a nonlinear effect. Indeed, Figures 18 and 19 provide further explanation by showing the first three FFT harmonics of STMA solution. Note that for ε=0.4, both higher harmonics show a significant presence in the shock region (approaching x=0.1), confirming the importance of nonlinear effects in this area.

Figure 14: Mean pressure contours for ε= 0.02 (STMA calculation).

Figure 15: Mean pressure contours for ε= 0.4 (STMA calculation).

The unsteady pressure predictions in the near to far fields surrounding the airfoil are illustrated in terms of the directivity contours for the predicted intensity of the unsteady pressure signal. Results shown in Figures 20-23 provide the directivity plots at radii R=1, 2, 3 and 4, respectively (as usual, all distances are normalized by the airfoil chord, and unsteady responses are scaled to the same input gust amplitude).

Figure 16: Airfoil RMS pressure on the suction and pressure sides: comparison of GUST3D solution with STMA predictions at ε=0.02, 0.2 and 0.4.

Figure 17: Base harmonic of airfoil pressure FFT on the suction and pressure sides: comparison of GUST3D RMS solution with STMA predictions at ε=0.02, 0.2 and 0.4.

Page 8: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

8

The issues of somewhat poor comparison with GUST3D predictions have been already discussed. It was also noted that at least in the linear limit, these results appear close to predictions in Ref. [4]. In the current calculations, a domain of smaller size was used, and as a result, some distortions due to the proximity of

Figure 18: Base, first and second harmonics of airfoil pressure FFT on the suction and pressure sides: comparison of GUST3D RMS solution with STMA predictions at ε=0.2.

Figure 19: Base, first and second harmonics of airfoil pressure FFT on the suction and pressure sides: comparison of GUST3D RMS solution with STMA predictions at ε=0.4.

the far field boundary and inherent moderate reflections may be noticeable at R=4. In order to further understand the nature of these distortions (particularly noticeable for ε=0.4), the directivity contours were calculated for the base and first FFT harmonics of the unsteady pressure signal.

Figure 20: Directivity of |p’2| for RMS pressure at R=1.

Figure 21: Directivity of |p’2| for RMS pressure at R=2.

Page 9: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

9

Clearly, the nonlinear effects increase significantly for ε=0.4, compared to low gust intensities (the axial spike in Figure 24d is actually due to the hydrodynamic pressure from the vortex which happens to locate at R=4). Still, the first harmonic remains one to two orders of magnitude lower than the base harmonic. Moreover, the base shows the same distortion of the upper lobe in Figure 24d observed for RMS pressure in Figure 23. Thus, although the higher harmonic does appear to extend in the direction of the lobe distortion (in Figure 25d), this effect may not be completely attributed to nonlinear generation of higher harmonics. Rather, this may also indicate the appearance of a non-compact acoustic source, or the problems with reflections from the far field boundary. To obtain a more clear picture of these phenomena, the contour plots of the unsteady pressure RMS and FFT are presented next.

Figure 22: Directivity of |p’2| for RMS pressure at R=3.

Figure 23: Directivity of |p’2| for RMS pressure at R=4.

Figure 24: Directivities of |p’2| for the base FFT harmonic, at R=1,2,3 and 4.

Figure 25: Directivities of |p’2| for the first FFT harmonic, at R=1,2,3 and 4.

Page 10: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

10

Qualitative Analysis of Nonlinear Response To further understand the qualitative behavior of the unsteady solution throughout the domain, the unsteady pressure contour plots are now examined. The linear response for ε=0.02 is observed in Figure 26. The contours show propagation of the acoustics away from the airfoil along with formation of the characteristic dipole-type radiation pattern observed in the previous directivity plots.

Figure 26: RMS of unsteady pressure for ε=0.02.

Figure 27: RMS of unsteady pressure for ε=0.2.

While the previous Figure contains no (at least apparent) indication of the nonlinear wake development, both cases of ε=0.2 and ε=0.4 (in Figures 27 and 28) show significant effects of nonlinear wake evolution, accompanied by strong hydrodynamic pressure pulsations. Enlarged in Figure 29 for ε=0.4, this region shows characteristic vortex roll-ups and pairings, indicative of nonlinear, inviscid wake instability. In general, acoustic radiation is known to develop from such nonlinear vortex interaction, but this is not evident from the current results. On the other

Figure 28: RMS of unsteady pressure for ε=0.4.

Figure 29: RMS of unsteady pressure for ε=0.4: airfoil and wake regions.

Page 11: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

11

hand, the presence of higher harmonics of unsteady pressure response is confirmed by comparing the base and the first FFT harmonics of the signal in Figures 30 and 31, for ε=0.4. The nonlinearity, exhibited by the presence of the first pressure harmonic, is clearly localized in the wake region, but it is also seen in the direction where the directivity lobe has been observed to develop a significant distortion (in Figure 23). Hence, at least some nonlinear effect appears present in this case. Finally, in order to quantify and localize effects from nonlinear gust-airfoil interaction, the difference of the appropriately scaled (by the corresponding amplitudes, for a unit input) base frequency and RMS of unsteady, linear pressure

Figure 30: FFT of unsteady pressure, base harmonic. ε=0.4.

Figure 31: FFT of unsteady pressure, first harmonic. ε=0.4.

response are shown in Figures 32-34 for ε=0.02, 0.2 and 0.4. Note that zero difference in this case is an indication of a purely linear response. Interestingly enough, some evidence of wake nonlinearity appears present even at ε=0.02, but the deviation from the linear behavior is very small (of the order of 1.e-4). On the contrary, the nonlinearities are clearly present for higher gust intensities. Those are still mostly localized in the wake, but are also again seen in the region above the airfoil. The relative amplitude of these nonlinearities appears to increase from about 1.e-3 for ε=0.2, to approximately 1.e-2 for ε=0.4, which confirms the previous observations.

Figure 32: Difference of base frequency and linear-response RMS of unsteady pressure, ε=0.02.

Figure 33: Difference of base harmonic and linear-response RMS of unsteady pressure, ε=0.2.

Page 12: [American Institute of Aeronautics and Astronautics 10th AIAA/CEAS Aeroacoustics Conference - Manchester, GREAT BRITAIN ()] 10th AIAA/CEAS Aeroacoustics Conference - STMA Study of

American Institute of Aeronautics and Astronautics

12

Figure 34: Difference of base harmonic and linear-response RMS of unsteady pressure, ε=0.4.

CONCLUSIONS The paper examined application of the Space-Time Mapping Analysis (STMA) method to predicting the nonlinear unsteady airfoil response to an impinging, finite-amplitude gust. By treating the time coordinate identically to the space directions, the STMA method essentially solved the unsteady 2D interaction problem as a steady-state 3D one. A high-order discretization scheme was applied to achieve time-accurate predictions of both unsteady aerodynamic and aeroacoustic responses. The results for the 2D small-amplitude gust interaction with a cambered, thick Joukowski airfoil at an angle of attack were calculated for the mean flow Mach number M=0.5 and gust reduced frequencies k=0.1, 1.0 and 2.0. For the case of k=1.0, solutions for a high-intensity gust-airfoil interaction were obtained, for gust amplitudes ranging from ε=0.02 to ε=0.4. To validate numerical computations, a comparison was conducted with the frequency-domain linear solver GUST3D, and another time marching solver. An excellent agreement was observed for the mean and RMS pressure predictions along the airfoil surface between all solutions. The far field acoustic directivities also showed an excellent comparison with the time marching solver. The results obtained from nonlinear high-amplitude gust calculations identified zones in the computational

domain where nonlinear response effects appeared most significant. Such zones have been detected primarily in the wake region, but were also observed in the region above the airfoil.

REFERENCES

1. Hixon, R., “Space-Time Mapping Analysis for the Accurate Calculation of Complex Unsteady Flows,” AIAA Paper 2003-3205, May 2003.

2. Scott, J.R. and Atassi, H.M., “A Finite-Difference, Frequency-Domain Numerical Scheme for the Solution of the Gust-Response Problem,” Journal of Computational Physics, Vol. 119, pp.75-93, 1995.

3. Golubev, V.V., Mankbadi, R.R. and Hixon, R., “Simulation of Airfoil Response to Impinging Gust Using High-Order Prefactored Compact Code,” to appear in Proceedings of the 4th CAA Workshop on Benchmark Problems, Cleveland, OH, October 2003.

4. Golubev, V.V., Mankbadi, R.R and Scott, J.R., “Numerical Inviscid Analysis of Nonlinear Airfoil Response to Impinging High-Intensity High-Frequency Gust,” AIAA Paper 2004-3002, 10th CEAS/AIAA Aeroacoustics Conference, Manchester, May 2004.

5. Tam, C. K. W. and Webb, J. C., “Dispersion-Relation-Preserving Finite-Difference Schemes for Computational Acoustics,” Journal of Computational Physics, Vol. 107, 1993, pp. 262-281.

6. Kennedy, C. A. and Carpenter, M. H., “Several New Numerical Methods for Compressible Shear-Layer Simulations,” Applied Numerical Mathematics, Vol. 14, 1994, pp. 397-433.

7. Jameson, A., Schmidt, W., and Turkel, E., “Numerical Solutions of the Euler Equations by Finite-Volume Methods Using Runge-Kutta Time-Stepping Schemes,” AIAA Paper 1981-1259, June 1981.

8. GridPro™, Program Development Corporation, White Plains, NY.

9. Golubev, V.V. and Rohde, A., “Application of Space-Time Mapping Analysis Method to Unsteady Nonlinear Gust-Airfoil Interaction Problem,” AIAA Paper 2003-3693, June 2003.