Вы находитесь на странице: 1из 18

ARTICLE IN PRESS

Progress in Aerospace Sciences 44 (2008) 4865 www.elsevier.com/locate/paerosci

Unsteady ow phenomena associated with leading-edge vortices


C. Breitsamter
Institute of Aerodynamics, Technische Universitat Munchen, BoltzmannstraX e 15, 85748 Garching, Germany

Abstract This paper presents selected results from extensive experimental investigations on turbulent ow elds and unsteady surface pressures caused by leading-edge vortices, in particular, for vortex breakdown ow. Such turbulent ows may cause severe dynamic aeroelastic problems like wing and/or n buffeting on ghter-type aircraft. The wind tunnel models used include a generic delta wing as well as a detailed aircraft conguration of canard-delta wing type. The turbulent ow structures are analyzed by root-mean-square and spectral distributions of velocity and pressure uctuations. Downstream of bursting local maxima of velocity uctuations occur in a limited radial range around the vortex center. The corresponding spectra exhibit signicant peaks indicating that turbulent kinetic energy is channeled into a narrow band. These quasi-periodic velocity oscillations arise from a helical mode instability of the breakdown ow. Due to vortex bursting there is a characteristic increase in surface pressure uctuations with increasing angle of attack, especially when the burst location moves closer to the apex. The pressure uctuations also show dominant frequencies corresponding to those of the velocity uctuations. Using the measured ow eld data, scaling parameters are derived for design purposes. It is shown that a frequency parameter based on the local semi-span and the sinus of angle of attack can be used to estimate the frequencies of dynamic loads evoked by vortex bursting. r 2007 Elsevier Ltd. All rights reserved.

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Models and experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Delta wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Delta-canard conguration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Test technique and test conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1. Wind tunnel facilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.2. Test methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.3. Test conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flow physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Leading-edge vortex ow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Turbulent ow elds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Surface pressure uctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. Fin buffet and buffeting characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Scaling parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Turbulence intensities and spectral densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Dominant frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 51 51 51 52 52 52 54 54 54 57 58 58 61 61 61 64

3.

4.

Abbreviations: CLV, canard leading-edge vortex; CTV, canard trailing-edge vortex; LEX, leading-edge extension; rms, root-mean-square; WLV, wing leading-edge vortex. Tel.: +49 89 28916137; fax: +49 89 28916139. E-mail address: Christian.Breitsamter@aer.mw.tum.de 0376-0421/$ - see front matter r 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.paerosci.2007.10.002

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 49

Nomenclature b CL CP(t) CP C 0P ^ CP span [m] lift coefcient pressure coefcient [p(t)pN]/qN mean (time averaged) pressure coefcient uctuation part of CP amplitude spectrum of pressure coefcient, p 2S C P DkU 1 =l m q 2 C Prms rms value of C 0P , C 0 P cr wing root chord [m] d thickness [m] f frequency [1/s] k reduced frequency, fl m =U 1 lC characteristic length [m] lm wing mean aerodynamic chord [m] p(t), pN pressure, ambient pressure [Pa] qN free stream dynamic pressure [Pa] Rel m Reynolds number, U 1 l m =n S CP pressure spectral density [s] Su0i power spectral densities of velocity uctuations u0i N S u0 power spectral density of u0 normalized with DkU 1 =u0 2 l m s, sl semi-span, local semi-span [m] t time [s] UC characteristic velocity [m/s]

free stream velocity [m/s] UN u, v, w axial (stream-wise), lateral and vertical velocity [m/s] u; v; w axial (stream-wise), lateral and vertical mean (time averaged) velocity [m/s] u0 , v0 , w0 uctuation part of u, v, w; u0i [m/s] p urms, vrms, wrms rms velocities of u0 , v0 , w0 ; urms u0 2 [m/s] Y, Z non-dimensional coordinates in lateral and vertical direction, Y y/s, Z z/s x, y, z Cartesian coordinates a angle of attack, deg L aspect ratio l taper ratio l wave length [m] n kinematic viscosity [m2/s] j leading-edge sweep [deg] r density [kg/m3] Subscripts B C dom F TE W bursting, burst location canard dominant n trailing-edge wing

1. Introduction Modern ghter aircraft are subject to high angle of attack maneuvers extending the ight envelope to the stall and post-stall regime [1]. Slender wing geometries, e.g. delta wing geometries, strakes, and leading-edge extensions (LEXs), respectively, are used to generate strong large-scale vortices shed at the highly swept leadingedges [2]. The leading-edge vortices improve signicantly the high angle of attack performance because of additional lift and an increase in maximum angle of attack. Time scales can be attributed to several unsteady ow phenomena in comparison to time scales associated with ow changes due to maneuvers [3] (Fig. 1). During maneuvers wing areas inuenced by ow separation or vortex bursting are shifted upstream or downstream at larger time scales while shorter time scales corresponding to unsteady ow phenomena, like shear layer instabilities, instability of the breakdown ow or wake instabilities, may be of great importance for dynamic aeroelastic effects [3,4]. At high angle of attack, the phenomenon of leading-edge vortex bursting over the wing planform is of specic interest [5,6]. The transition from stable to unstable core ow, evident by the rapid change in the axial velocity proles from jet- to wake-type, leads to extremely

high turbulence intensities at the breakdown position and to increased turbulence levels further downstream [3]. Hence, the buffet excitation level increases strongly above a certain angle of attack, and wing and n normal force spectra may exhibit narrow-band peaked distributions (Fig. 2). Such unsteady aerodynamic loads often excite the vertical tail structure or even the wing structure in their natural frequencies, resulting in increased fatigue loads, reduced service life and raised maintenance costs [4,7]. For example, the n buffeting problem plagues twin-n congurations (F-15, F/A-18), but single-n aircraft are also affected [4,7]. Therefore, comprehensive research programs have been undertaken aimed at understanding the buffet loads and reducing the structural response. The related vortex ow features are carefully analyzed using wind tunnel tests on small- and full-scale models [710], supplemented by ight tests [7,11], and detailed numerical ow simulations [12,13]. In addition, design methods have been developed to describe the n buffet environment and to predict buffet loads for use in aircraft design [14,15]. The buffet loads do not only decrease the fatigue life of the airframe, but may, in turn, limit the angle of attack envelope of the aircraft. To counter buffeting problems, several methods have been suggested. They deal with alterations of the structural properties

50

C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

Fig. 1. Unsteady ow phenomena on delta wing congurations: ow features associated with time scales of tUN/crb1 (a) and ow features associated with time scales of tUN/crp1 (b).

Fig. 2. Characteristics of vortex induced n buffeting.

like stiffness and damping [16], aerodynamic modications for passive or active control of vortex trajectories to avoid a direct impact of the burst vortex ow [17] and methods of active vibration control [1822]. Using active control, the structural dynamic loads can be reduced aimed both to increase the service live and to enhance the maneuverability by extending the angle of attack envelope. To improve the knowledge on the ow physics associated with high angle of attack buffet problems caused by

unsteady vortex ows, extensive experiments have been conducted at the Institute of Aerodynamics (AER) of the Technische Universitat Munchen (TUM) [3,15,22,23]. The studies focus on the low-speed ow environment of generic delta wing models, as well as of a modern ghter aircraft model. The turbulent ow structure is analyzed by the spatial and temporal characteristics of the time-dependent ow velocities supplemented by unsteady surface pressures. This paper gives an overview of the main results characterizing the unsteady ow phenomena by turbulence

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 51

intensities, spectral densities and dominant reduced frequencies. Such data are of great importance for ghter aircraft design.

2. Models and experiment 2.1. Delta wing The delta wing model is a sharp-edged carbon ber model with a root chord of 0.670 m and a span of 0.335 m (Fig. 3). The leading-edge sweep is 761 corresponding to an aspect ratio of 1. Fast-response differential pressure transducers are embedded in the right part of the upper surface located at root chord positions of x/cr 0.3 (9 transd.), x/cr 0.5 (9 transd.), x/cr 0.7 (11 transd.), and x/cr 0.8 (10 transd.). At 90% root chord 25 transducers are positioned along the entire local span to prove symmetric ow conditions. 2.2. Delta-canard conguration The delta-canard model consists of nose section, front fuselage including rotatable canards and a single place canopy, center fuselage with delta-wing section and a through-ow air intake underneath, and rear fuselage including nozzle section and the vertical tail (n) (Fig. 4). All model parts are made of stainless steel. The wing sweep is 501, the canard sweep 451 and the span of the model is 0.740 m. For the model tested, the leading- and

Fig. 3. Geometry of generic delta wing conguration.

Fig. 4. Geometry and wind tunnel model of ghter aircraft conguration: geometry of delta-canard conguration (a) and views of 1/15-scale delta-canard model mounted in test section of TUM-AER low-speed wind tunnel facility B (b).

ARTICLE IN PRESS
52 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

trailing-edge aps as well as the canard were set to 01. The n is instrumented with 18 differential pressure transducers at nine positions directly opposite each other on each surface. 2.3. Test technique and test conditions 2.3.1. Wind tunnel facilities The investigations are carried out in the Gottingen type low-speed wind tunnel facilities A and B of the AER of the TUM. Dimensions of the open test sections (height width length) are 1.8 m 2.4 m 4.8 m (W/T A) and 1.2 m 1.55 m 2.8 m (W/T B), respectively. Maximum usable velocity is 65 and 60 m/s, respectively. The test section ow was carefully inspected and calibrated, documenting a turbulence level less than 0.4% and uncertainties in the spatial and temporal mean velocity distributions of less than 0.067%. The blockage at maximum angle of attack is below 6%. The models are

sting mounted using a computer-controlled three-axis model support (Fig. 4b). 2.3.2. Test methods Advanced hot-wire anemometry with miniature dualand triple-sensor probes was used to measure the timedependent velocity components. The sensors consist of 5-mm-diameter platinum-plated tungsten wires giving a length/diameter ratio of 250. The measuring volume formed by the wires is approximately 1 mm in diameter. The wires are arranged perpendicular to each other to achieve best angular resolution [3,24]. An additional temperature probe is employed to correct anemometer output voltages if ambient ow temperature varies. The hot-wire probes were operated by a multi-channel constant-temperature anemometer system. By means of its signal conditioner modules, bridge output voltages were typically low-pass ltered at 1000 Hz before digitization and amplied for optimal signal level. The signals were

Fig. 5. Delta wing vortex formation: main delta wing ow features (a) and vortex bursting characteristics (b).

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 53

then digitized with 16-bit precision through the sixteenchannel simultaneous-sampling A/D converter of a PC high-speed board. The sampling rate for each channel was set to 3000 Hz. The sampling time was 6.4 s, resulting in a sample block of 19,200 values for each survey point. The sampling parameters were achieved by preliminary tests to ensure that all signicant ow eld phenomena are detected. Each digitized and temperature-corrected voltage triple was converted to calculate the time-dependent velocity components u, v, and w. The numerical method used is based on look-up tables derived from the full velocity and ow angle calibration of the sensors [3,24].

Statistical accuracy of the calculated quantities was considered as well. Random error calculations gave accuracies of 0.5%, 2%, and 3.5% for the mean and standard deviation and spectral density estimation, respectively [24]. Regarding the susceptibility of vortex structures to intrusive measurements, it was found that the presence of the probe has no markable inuence on the vortex structures of interest [3]. The uctuating surface pressures were recorded by transducers of active differential type giving a best signal resolution of about 2 Pa. The sensors are connected to a data acquisition system with 8 modules each tted with 16

Fig. 6. Typical delta wing lift polar and ow eld topology as function of angle of attack; jW 761, UN 37 m/s, Relm 1.07 106.

ARTICLE IN PRESS
54 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

differential ampliers, low-pass lters of Butterworth type, and sample and hold units linked to a 14-bit A/D converter. Thus, 128 signal inputs can be treated simultaneously. Here, the sampling rate for each channel was set to 2000 Hz and the low-pass lter frequency to 256 Hz. The sampling time is 30 s, giving a sample block of 60,000 values for each transducer. Statistical errors are below 0.5%. 2.3.3. Test conditions The delta wing tests were made at a free stream reference velocity of UN 37 m/s, giving a Reynolds number of Relm 1.07 106. The angle of attack is varied in the range 01pap601. The results presented focus on symmetric free stream conditions. The velocity measurements were performed in planes normal to the wing surface located directly above the pressure transducer sections. The measurement points are evenly spaced in lateral and vertical direction with a relative distance of 0.045 based on the local wing semi-span. The conditions of the delta-canard investigations correspond to those of the delta wing studies: UN 40 m/s, Relm 0.97 106, a 01301. The relative spacing of oweld survey points is 0.027 laterally and 0.042 vertically, also based on the local wing semi-span. At all tests, turbulent boundary layers were present at wing and control surfaces, proven by shearstress sensitive liquid crystal measurements [3]. 3. Flow physics 3.1. Leading-edge vortex ow Delta wing planforms representing lifting surfaces with highly swept leading-edges and low aspect ratios have been investigated intensively over the last 50 years [2,5,6]. Here, the interest is particularly on low-speed cases. Therefore, compressibility effects are not addressed. The dominating ow eld characteristics are given by the evolution and development of two large-scale vortices shed at the delta wing leading-edges (Fig. 5a). Vortex formation starts already at low angles of attack developing from the wing rear part to the apex. The separating shear layers from the wing upper and lower surface roll up by selfinduction to form a vortex, which is positioned over the wing. This primary vortex is fully developed when vorticity feeding exists over the whole leading-edge. The vortex core shows high axial velocities, low static pressures and high dissipation, i.e. lower total pressures, in the sub-core area due to the steep gradient in the cross ow components. The leading-edge vortices increase the velocities on the wing upper surface. This velocity increase leads to a high suction level, with the local pressure minima indicating the track of the vortex axis on the wing surface. Consequently, leadingedge vortices in a fully developed, stable stage create additional lift and an increase in maximum angle of attack, improving signicantly maneuver capabilities of highagility aircraft.

Considering a chord-wise station, the pressure increases from the suction peak induced by the primary vortex to the leading-edge, resulting in a severe lateral pressure gradient. Typically, a further separation occurs, forming a secondary, counter-rotating vortex, the evolution of which depends strongly on the presence of a laminar or turbulent boundary layer [25]. Further, leading-edge vortices are subject to breakdown at high angles of attack (Fig. 5b). Vortex breakdown is caused by the stagnation of the axial core ow due to the increase of the adverse pressure gradient when raising the angle of attack [5,6,26]. Therefore, the core expands rapidly accompanied by high velocity uctuations [3]. Delta wing research activities often focus on a sharp leading-edge because primary separation is xed and the leading-edge vortex development is only little inuenced by Reynolds number effects. A blunt leading-edge complicates the vortex aerodynamics as the position of the separation line is free to move, determined by the pressure gradient and the boundary layer development [25]. Thus, leadingedge radius, angle of attack and Reynolds number are the main parameters adjusting the onset of vortex evolution as well as position and strength of the primary vortex. For the sharp leading-edge case, the angle of attack is the main parameter only [27].

Fig. 7. Delta wing vortex ow stages as function of wing sweep jW and angle of attack a.

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 55

In this context, Fig. 6 shows the lift polar of the investigated 761 delta wing conguration supplemented by corresponding ow elds. The vortex formation is schematically represented and the vector eld of the cross-ow velocities is shown at 90% wing root chord for angles of attack of a 12.51, 251, 301 and 351. The velocity eld clearly indicates that the vortex strength increases with increasing angle of attack. The two primary vortices move inboard and upward and their cross sections become enlarged, cp. a 12.51 and 251. Above a certain angle of

attack, the primary vortices move purely upward due to the mutual displacement effect of the two vortices. The vortexinduced velocities create high suction on the wing, leading to a nonlinear increase in the lift coefcient. Raising the angle of attack further changes the vortex core structure, a 301. This change reveals itself as a sudden expansion of the vortex core ow, known as vortex bursting or vortex breakdown. The expansion of the vortex core ow is accompanied by a substantial decrease of the vortexinduced velocities. Bursting starts over the wing at the

Fig. 8. Typical delta wing turbulence intensity elds for fully developed and burst leading-edge vortices; jW 761, UN 37 m/s, Relm 1.07 106.

ARTICLE IN PRESS
56 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

trailing-edge and the breakdown position moves forward with increasing angle of attack, a 351. Thus, the wing upper surface is more and more affected by the breakdown ow how the suction level becomes diminished and the lift coefcient decreases. For Reynolds numbers of Re4104 and typical swirl numbers of delta wing leading-edge vortices vortex breakdown is of the spiral type. A correlation between wing sweep (fW 451851) and angle of attack (a 01401) is shown in Fig. 7, concentrating on thin wings with sharp leading-edge. A rst range can be assigned to the evolution of the primary vortex starting from the wing rear part to the apex. Consequently, the primary vortex is only present within the rearward range of the wing and is not completely rolled up yet. The dotted line indicates the vortex axis. The second range marks that of the fully developed vortex, the axis of which moves inboard and upward with increasing angle of attack. The transition from range 1 to range 2 depends on the

boundary layer condition, namely laminar or turbulent. For a given wing sweep, the laminar boundary layer creates the fully developed vortex at a smaller angle of attack relative to the turbulent case. The next range marks the case of the span-wise xed vortex. Here, the vortex axis is only shifted upward with increasing angle of attack. The transition from range 2 to range 3 depends like before on whether a laminar or turbulent boundary layer is present. Furthermore, it can be stated that the span-wise xed vortex develops only for a wing sweep larger than jWE651. The aerodynamic benets resulting from leading-edge vortices become limited when vortex bursting affects markedly the wing ow. At moderate wing sweep, vortex bursting exists over a large incidence range until the maximum angle of attack is reached. These ranges of leading-edge vortex conditions determine signicantly the aircraft maneuverability at moderate and high angles of attack.

Fig. 9. Characteristic delta wing turbulence intensity distributions and power spectral densities for the leading-edge vortex breakdown ow eld; jW 761, Relm 1.07 106.

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 57

Regarding numerical simulations, great success has been achieved in the development and application of high delity computational uid dynamics (CFD) methods in the last 10 years [2830]. Unsteady Reynolds averaged Navier-Stokes (URANS) methods are available including a variety of turbulence models based on algebraic up to Reynolds stress transport equations. Also, methods for detached eddy simulations (DES) are formulated as a combination of a large eddy simulation (LES) to model separated ow dominated by large-scale structures in the outer domain and a turbulence model to calculate ow quantities in the wall-bounded domain. Even the upper wing surface pressure distribution is very sensitive to the correct modeling of viscous effects on the wing as well as in the rolled-up shear layers. The calculation and analysis of unsteady loads is even a more challenging problem, which

needs the correct representation of the turbulent and timedependent ow eld features [22]. 3.2. Turbulent ow elds The turbulent ow elds are analyzed using root-meansquare (rms) values of the velocity uctuations referred to the free stream velocity UN. Contours of axial rms velocities are shown for the 761 delta wing conguration, plotted again for the cross-ow plane located at 90% wing root chord and angles of attack of a 12.51, 251, 301 and 351 (Fig. 8, cp. Fig. 6). The fully developed vortices are indicated by regions of moderate turbulence intensities including local maxima of 5%2%, a 12.51 and 251. These regions clearly depict the shear layers separating from the leading-edge and

Fig. 10. Delta wing rms pressures as function of angle of attack considering especially cases of leading-edge vortex bursting over the wing; jW 761, Relm 1.07 106.

ARTICLE IN PRESS
58 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

rolling up to form the rotational core of the primary vortices. The (viscous) sub-core of the primary vortex is indicated by a circle-like region of local rms maxima of about 10%12%. Also, the secondary vortices are marked by increased turbulence intensities. The areas of local rms maxima detected in the shear layers represent the formation of coherent sub-structures associated with a shear layer instability of the KelvinHelmholtz type [3] (cp. Fig. 1). Because of vortex bursting at x/cr 0.86, the ow pattern of the vortex core is signicantly changed at a 301. Next to the original jet-like core, a region of strong ow deceleration occurs. This region of retarded axial core ow is caused by the adverse pressure gradient arising at high angle of attack. The corresponding steep velocity gradients and the rapid change from jet- to wakelike core ow evokes an overall maximum in turbulence intensity at the vortex center. With increasing angle of attack, a 351, vortex bursting occurs much more upstream, namely at x/cr 0.49. Downstream, the region of maximum turbulence intensity expands rapidly, which is evident at 90% wing root chord. The associated local turbulence maxima of 512%, are located in a limited radial range around the burst vortex core. This area corresponds to the points of inection in the radial proles of the retarded axial core ow. This development of rms patterns with increasing angle of attack for a xed chordwise station can also be determined when xing the angle of attack and moving downstream from the wing apex to the trailing-edge (Fig. 9). Analyzing the spectral content of the velocity uctuations it is shown that the breakdown ow exhibits a signicant spectral peak, indicating that turbulent kinetic energy is channeled into a narrow band. The frequency related to this spectral peak is named dominant frequency. The energy concentration in a limited frequency range is linked to a specic instability mechanism called helical mode instability of the breakdown ow [3,9]. Hence, quasi-periodic aerodynamic loads result, which may strongly excite structural modes. 3.3. Surface pressure uctuations As shown before, the impingement of burst leading-edge vortices is a source of buffet excitation on an aircraft experienced on the wing surface or on other surfaces such as the n. The buffet excitation is substantiated by the corresponding pressure uctuations measured on the delta wing surface (Fig. 10). The rms pressures are plotted as a function of angle of attack for the symmetry location at 90% wing root chord. In addition, patterns of rms surface pressures are shown for a 351, 401, and 451. If vortex breakdown takes place in the rear part of the wing the rms surface pressure distribution is less affected. With the breakdown position moving upstream the pressure uctuation intensities increase strongly, especially beneath the vortex axis, where the distance to the region of maximum ow eld turbulence is the smallest. A noteworthy situation

Fig. 11. Flow eld characteristics of delta-canard conguration at a 151; UN 40 m/s, Relm 0.97 106: cross-ow velocity vector elds at 40% and 100% root chord position (a) and topology of main vortex systems (b).

exists at a 451, when the surface pressure uctuations reach an absolute maximum with rms levels above 30%. A further increase in angle of attack shifts the breakdown position closer to the apex and the rms pressures decrease. This decrease is due to the detachment of the vortex axis reducing the impingement of the highly uctuating ow eld on the wing surface. 3.4. Fin buffet and buffeting characteristics At TUM-AER comprehensive studies on the vortex ow elds associated with wing and n buffeting have been conducted on generic models as well as on the EF-2000 type delta canard conguration [3,15,23,24] (Fig. 4). The overall ow eld development is shown in Fig. 11 including a schematic representation of the main vortex systems based on laser light sheet tests. The dominating vortices are the wing leading-edge vortices

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 59

Fig. 12. Lateral rms velocities vrms/UN measured in the n region and in a plane normal to the n for various angles of attack; UN 40 m/s, Relm 0.97 106: rms values for different vertical n stations Z (a) and cross-ow rms patterns and velocity vectors at a 201 (b); at a 251 (c); and a 301 (d).

(WLVs) and the canard leading-edge vortices (CLVs) and canard trailing-edge vortices (CTVs) inuencing the ow eld in the n region. The impact of the ow eld on the n structure can be characterized by the lateral rms velocities. Summarizing, the rms values for different vertical n stations are shown as a function of angle of attack in Fig. 12a. The magnitude of the rms values in the midsection depends on the development of the vortex ow structure. This is depicted by the schematics of Fig. 12a, which are based on the rms velocity patterns in planes normal to the n surface (Figs. 12bd). At moderate angles of attack the area of the center-line n is only little affected by the regions of highly turbulent ow attributed to the burst wing leading-edge vortices and the canard leadingand trailing-edge vortices (Fig. 12b). Above a 251, the

lateral turbulence intensities in the n region increase signicantly with increasing angle of attack as the burst WLV expand and approach the midsection (Fig. 12c). Also the canard leading- and trailing-edge vortices come close to the midsection. In particular, the n ow is inuenced by induction effects arising from the WLV sheets, which are the loci of maximum turbulence intensity. The interaction between the WLVs and the canard vortices (CLVs and CTVs) leads to local rms maxima near the n tip (Fig. 12d). The unsteady ow eld induces pressure uctuations on the n. The surface-averaged rms values of the uctuations in the pressure coefcient are plotted as a function of angle of attack dening the buffet situation (Fig. 13). Representing the trend in the lateral turbulence intensities, the rms pressure coefcient increases signicantly above a 201,

ARTICLE IN PRESS
60 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

Fig. 13. Surface-averaged n rms pressures as function of angle of attack; UN 40 m/s, Relm 0.97 106.

^ Fig. 14. Amplitude pressure spectra C P taken at n station P13 for all angles of attack tested; UN 40 m/s, Relm 0.97 106.

reaching a value of about 8% at a maximum angle of attack of aE301. The severe increase in the rms pressure above a certain incidence is a characteristic feature of the n buffet phenomenon (cp. Fig. 2). The amplitude spectra of the uctuating pressure coefcient, calculated from the signal taken at sensor station P13, are shown in Fig. 14 for all angles of attack tested. Above aE221, spectral peaks can be identied in the range of reduced frequencies of k 0.80.6 (k f lm/UN). The helical mode instability of the burst WLVs starts to affect the n pressure eld and the narrow-band amplitude increases strongly from a 241 to a 31.21. This energy peak is called buffet peak. Hence, the narrow-band concentration of turbulent kinetic energy may result in

strong excitation of structural modes. It can be further detected that the reduced frequencies associated with the buffet peak, i.e. the dominant frequencies, are shifted to lower values at higher angle of attack. Such pressure distributions create the buffeting or structural response to the buffet. The resulting n buffeting mainly consists of a response in the rst bending and torsion mode. Thus, the uidsstructure interaction of vortex breakdown with a n involves the following phenomena (Fig. 15): the time-averaged breakdown location depending on the adverse pressure gradient set by the recompression at the wing trailing-edge and/or by the blockage of the n, the helical mode instability of the breakdown ow, quasiperiodic oscillations of the breakdown location, distortion

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 61

structural dynamic properties. As low-speed high-angle-ofattack maneuvers are of primary interest, compressibility effects are not taken into account. 4.1. Turbulence intensities and spectral densities Regarding the n buffet problem, velocity uctuations are measured in the n region of the delta-canard conguration (midsection: Y 0.0, Z 0.21) for different angles of attack, varying the free stream velocities in the low-speed environment (Fig. 16). It is shown that the rms velocities based on the free stream velocity are constant over the considered velocity range. Hence, the rms velocities can be scaled by referring to the free stream velocity UN. Similarly, rms pressures are scaled by referring to dynamic free stream pressure qN. Further, the power spectral density values of the velocity uctuations can be made non-dimensional with rms velocities and a characteristic length lC and free stream velocity UN as reference parameters. Thus, the nondimensional rms values q u0 2 i (1) U1 and the non-dimensional spectral densities Su0i Dk U 1 u0 2 l m
i

(2)

can be used as appropriate scaling quantities regarding turbulence effects. 4.2. Dominant frequency The velocity uctuations measured in the n region of the delta-canard conguration at Y 0.0, Z 0.21 are characterized by concentrations of turbulent kinetic energy at specic frequencies. These narrow-band concentrations of the velocity uctuations in the n region are mainly evoked by induction effects of the uctuations associated with the approaching burst WLVs (see Figs. 9, 12 and 15). The corresponding dominant reduced frequencies kdom of the incident vortex and vortex splitting, unsteady ow separation at the n leading-edge, and possible coupling between the separated n ow and/or n elastic deections with oscillations of the breakdown location. Among these, the dominant phenomenon causing n buffeting is the quasi-periodic loading on the n due to the helical mode instability of the leading-edge vortex breakdown ow. 4. Scaling parameters The following parameters are of main importance for design issues. Especially the frequency parameter associated with dominant buffet loads is of specic interest for f dom l C , UC (3)

Fig. 15. Phenomena related to n buffeting: visualization of F-18 LEX vortices bursting in front of the ns (a) and uidstructures interaction of leading-edge vortex bursting and (elastic) n (b).

are scaled with lC=lm and UC=UN and plotted as a function of free stream velocity UN for different angles of attack (Fig. 17). The graph substantiates that the reduced frequency values do not change with free stream velocity UN, but decreases with angle of attack a. This shift in the dominant reduced frequency to lower values with increasing angle of attack was also found for the amplitude spectra of the n surface pressure uctuations (cp. Fig. 14). For further analysis, the dominant reduced frequency as a function of angle of attack is plotted in Fig. 18. The quasiperiodic velocity and induced surface pressure uctuations, respectively, result from the helical mode instability of the

ARTICLE IN PRESS
62 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

Fig. 16. Lateral turbulence intensity (rms velocity vrms/UN) measured at n station (xW/cr 1.13, Y 0.0, Z 0.21) of the delta-canard conguration as a function of free stream velocity UN for angles of attack of a 20.031.51; Relm 0.17 1061.22 106.

sinus of angle of attack and the co-tangent of the wing leading-edge sweep: k dom f dom l C f dom x cot jW sin2 a U 1 sin a UC f dom x cot jW sin a 0:28 0:025. U1 |{z}
kdom

Considering especially the n region of the delta-canard conguration at a downstream position of the wing trailing-edge, Eq. (4) is written with x cr, i.e., k dom
Fig. 17. Dominant reduced frequency kdom of the lateral velocity uctuations at n station (xW/cr 1.13, Y 0.0, Z 0.21) of the deltacanard conguration as function of free stream velocity UN for angles of attack of a 25.01, 28.01, and 31.51; Relm 0.17 1061.22 106.

f dom cr cot jW sin a 0:28 0:025. U1 |{z}


kdom

(5)

ow downstream of vortex breakdown. The burst vortex core expands with increasing angle of attack and, therefore, the wavelength of the instability mode becomes larger and the corresponding frequency smaller, fdom$1/l$1/sl. A universal frequency parameter k dom can be derived using appropriate scaling quantities (Fig. 19). Referring to velocity, the component normal to the leading-edge (UN sin a) has to be considered. The characteristic length scale lC must account for the vortex core expansion of the burst leading-edge vortex given approximately by the local half span ($sl x cot jW) and the shear layer distance ($sin2 a). Using these relations leads to a scaling with the

The application of this frequency parameter is shown in Fig. 19 for wing sweeps of jW 501 (delta-canard conguration, Fig. 3) and jW 761 (delta wing conguration, Fig. 4). Measurements on different congurations substantiate the validity of the derived frequency parameter [3,9,10,14,30]. Fig. 18 also demonstrates that this scaling groups the values of the dominant reduced frequencies taken in the n region indeed within a band of k 0:28 0:025. dom Thus, the frequency f dom indicating a dominant energy accumulation exciting structural modes can be derived for a certain conguration (jW) at a specic ight condition (UN, a): f dom 1 U1 0:28 0:025. cr cot jW sin a (6)

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 63

Fig. 18. Application of the buffet load frequency parameter kdom* to the n buffeting problem studied at the delta-canard conguration: schematic representation of dominant n buffeting phenomena (cf. Fig. 15) (a) and dominant reduced buffet frequency kdom as a function of angle of attack based on amplitude pressure spectra of n station P13 (cp. Fig. 14); UN 40 m/s, Relm 0.97 106 (b).

5. Conclusions and outlook Detailed experimental investigations on the turbulent ow elds associated with fully developed and with burst leading-edge vortices have been conducted on a 761 swept delta wing and on a ghter aircraft model of canard-delta wing type with a wing sweep of 501. The measurements include eld distributions of velocity uctuations based on advanced hot-wire anemometry and distributions of surface pressure uctuations obtained by unsteady pressure transducers. Spectral quantities are analyzed as well. Major results are as follows: (1) Fully developed leading-edge vortices are characterized by moderate turbulence levels for the rolled-up shear layers and the vortex core, ui,rmsE6%12%. Pressure uctuations induced on the shedding surface (wing) are low, cprms p3%. (2) The turbulence intensities rise signicantly when vortex breakdown takes place. At the burst position the velocity uctuations show an overall maximum near the vortex center, whereas in the breakdown wake the velocity uctuations show local maxima in a limited radial range (annular region), ui,rmsE18%28%. Thus, the buffet excitation on an airframe component increases strongly when the distance between the vortex

axis and the wing or n surface becomes smaller than one diameter of the burst vortex core. (3) Quasi-periodic velocity uctuations occur downstream of vortex bursting corresponding to a helical mode instability of the breakdown ow eld. The amplitudes of these narrow-band uctuations are the largest in the annual region of maximum turbulence intensities. These frequency-dependent velocity uctuations give also rise to coherent unsteady pressures, the frequency of which corresponds to that of the unsteady velocities. (4) A universal frequency parameter can be attributed to the dominant frequencies fdom for the velocity and pressure uctuations of the breakdown ow eld. This frequency parameter is based on a length scale for the lateral expansion of the burst vortex core and a characteristic velocity given by the wing leading-edge normal velocity component, resulting in the relation f dom cr =U 1 cot jW sin a 0:28 0:025: This reduced frequency can be used for design purposes to measure the dominant frequencies linked to dynamic aeroelastic problems, like vortex-induced wing and n buffeting, when varying wing sweep and ight conditions (velocity, angle of attack). (5) In the presence of leading-edge vortices, the helical mode instability of the breakdown ow at high angle of attack is the leading mechanism for inducing severe

ARTICLE IN PRESS
64 C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865

Fig. 19. Buffet load frequency parameter kdom* obtained by scaling the dominant reduced buffet frequency kdom with the sinus of angle of attack a and the cotangent of wing semi-span cot jW.

narrow-band unsteady loads (buffet) on aircraft wing and stabilizers. The turbulence quantities reported herein serve also as a database for numerical investigations as performed in the EC-funded research projects FLOMANIA (Flow Physics ModellingAn Integrated Apporach) and DESider (Detached Eddy Simulations for Industrial Aerodynamics), aimed at validating, improving and developing methods in the eld of hybrid RANS-LES methods. References
[1] Herbst WB. Future ghter technologies. AIAA J Aircraft 1980;17(8): 5616. [2] Hummel D. On the vortex formation over a slender wing at large angles of incidence. High angle of attack aerodynamics, AGARDCP247, Sandefjord, Norway, October 46, 1978. p. 15-117.

[3] Breitsamter C. Turbulente Stromungsstrukturen an Flugzeugkon gurationen mit Vorderkantenwirbeln. Dissertation, Technische Universitat Munchen, Herbert Utz Verlag Wissenschaft (Aerodynamik), 1997, ISBN 3-89675-201-4. [4] Luber W, Becker J, Sensburg O. The impact of dynamic loads on the design of military aircraft. Loads and requirements for military aircraft, AGARD-R-815, AGARD, Neuilly Sur Seine, France, 1996. p. 8-127. [5] Hummel D. Untersuchungen uber das Aufplatzen der Wirbel an schlanken Deltaugeln. Zeitschrift fur Flugwissenschaften und Weltraumforschung, Band 13, 1965. p. 15868. [6] Lambourne NC, Bryer DW. The bursting of leading edge vortices. Some observations and discussion of the phenomenon. ARC R M 3282 1962. [7] Lee BHK, Brown D, Zgela M, Poirel D. Wind tunnel investigations and ight tests of tail buffet on the CF-18 aircraft. Aircraft dynamic loads due to ow separation, AGARDCP483, Sorrento, Italy, April 16, 1990. p. 1-126. [8] Canbazoglu S, Lin JC, Wolfe S, Rockwell D. Buffeting of ns: distortion of incident vortex. AIAA J 1995;33(11):214450.

ARTICLE IN PRESS
C. Breitsamter / Progress in Aerospace Sciences 44 (2008) 4865 [9] Gursul I, Xie W. Buffeting ows over delta wings. AIAA J 1999;37(1):5865. [10] Meyn LA, James KD. Full-scale wind tunnel studies of F/A-18 tail buffet. AIAA J Aircraft 1996;33(3):58995. [11] Del Frate JH, Zuniga FA. In-ight ow eld analysis on the NASA F-18 high alpha research vehicle with comparisons to ground facility data. AIAA paper 90-0231, January 1990. [12] Rizk YM, Gee K. Unsteady simulation of viscous oweld around F-18 aircraft at large incidence. AIAA J Aircraft 1992;29(6): 77381. [13] Kandil OA, Sheta EF, Liu CH. Computation and validation of uid/structure twin tail buffet response. In: Euromech colloquium 349, simulation of structure uid interaction in aeronautics. Gottingen, Germany: German Aerospace Research Center; 1996. p. 15-115-10. [14] Ferman MA, Patel SR, Zimmermann NH, Gerstenkorn G. A unied approach to buffet response of ghter aircraft empennage. Aircraft dynamic loads due to ow separation, AGARDCP483, Sorrento, Italy, April 16, 1990. p. 2-118. [15] Breitsamter C, Laschka B. Fin buffet pressure evaluation based on measured oweld velocities. AIAA J Aircraft 1998;35(5): 80615. [16] Ferman MA, Liguore SL, Smith CM, Colvin BJ. Composite Exoskin doubler extends F-15 vertical fatigue life. AIAA paper 93-1341, April 1993. [17] Hebbar SK, Platzer MF, Frink WD. Effect of leading-edge extension fences on the vortex wake of an F/A18 Model. AIAA J Aircraft 1995;32(3):6802. [18] Ashley H, Rock SM, Digumarthi RV, Chaney K, Eggers Jr AJ. Active control for n buffet allevation. US Air Force Wright Lab., WL-TR-93-3099, Wright Patterson AFB, OH, January 1994. [19] Becker J, Luber W. Comparison of Piezoelectric and aerodynamic systems for aircraft vibration alleviation. In: SPIE 5th Annual symposium on smart structures and materials, conference paper 332604, San Diego, CA, March 1998. 65 [20] Galea SC, Ryall TG, Henderson DA, Moses RW, White EV, Zimcik DG. Next generation active buffet suppression system. AIAA paper 2003-2905, July 2003. [21] Sheta EF. Alleviation of vertical tail buffeting of F/A18 aircraft. AIAA J Aircraft 2004;41(2):32230. [22] Breitsamter C. Aerodynamic active control of n-buffet load alleviation. AIAA J Aircraft 2005;42(5):125263. [23] Breitsamter C, Laschka B. Turbulent ow structure associated with vortex-induced n buffeting. AIAA J Aircraft 1994;31(4):77381. [24] Breitsamter C, Laschka B. Velocity measurements with hot-wires in a vortex-dominated ow eld. Wall interference, support interference and ow eld measurements, AGARDCP535, Brussels, Belgium, October 47, 1993. p. 11-113. [25] Hummel D. Effects of boundary layer formation on the vortical ow above slender delta wings. Enhancement of NATO military ight vehicle performance by management of interacting boundary layer transition and separation. Prag, Czech Republic, October 47, 2004. [26] Nelson RC, Visser KD. Breaking down the delta wing vortex. Vortex ow aerodynamics, AGARDCP494, Scheveningen, The Netherlands, October 14, 1991. p. 21-115. [27] Furman A, Breitsamter C. Investigation of ow phenomena on generic delta wing. In: ICAS proceedings, 25th International congress of the aeronautical sciences. Hamburg, Germany, September 1318, 2006. p. 312-116. [28] Rai P, Finley DB, Ghaffari F. An assessment of CFD effectiveness for vortexow simulation to meet preliminary design needs, Vortex ow and high angles of attack. Loen, Norway, May 711, 2001. [29] Gurr A, Rieger H, Breitsamter C, Thiele F. Detached-eddy simulation of the delta wing of a generic aircraft conguration. Notes on numerical uid mechanics, new results in numerical and experimental uid mechanics, V. Contributions to the 14th AG STAB/DGLR symposium, Bremen, 2004 (NNFM), vol. 92. Berlin: Springer Verlag; 2006. [30] AVT113/VFE2: /http://cawap-prism.larc.nasa.gov/CAWAPI/ DELTAWING/S.

Вам также может понравиться