Sie sind auf Seite 1von 15

Control of Wing Vortices

I. Gursul, E. Vardaki, P. Margaris, and Z. Wang


Department of Mechanical Engineering University of Bath Bath, BA2 7AY, United Kingdom

Summary
Vortex control concepts employed for slender, nonslender and high aspect ratio wings were reviewed. For slender delta wings, control of vortex breakdown has been the most important objective, which is achieved by modifications to swirl level and pressure gradient. Delay of vortex breakdown with the use of control surfaces, blowing, suction, high-frequency and low-frequency excitation, and feedback control was reviewed. For nonslender delta wings, flow reattachment is the most important aspect for flow control methods. For high aspect ratio wings, vortex control concepts are diverse, ranging from drag reduction to attenuation of wake hazard and noise, which can be achieved by modifications to the vortex location, strength, and structure, and generation of multiple vortices.

1 Introduction
Controlling vortical flows over wings may have various benefits, such as enhancement of lift force, generation of forces and moments for flight control, attenuation of buffeting, reduction of drag, and attenuation of noise due to vortex/blade interaction. Control methods include manipulation of one or more of the following flow phenomena: flow separation from the wing, separated shear layer, vortex formation, flow reattachment on the wing surface, and vortex breakdown. The occurrence and relative importance of these phenomena strongly depend on the wing sweep angle. For delta wings, flow reattachment and vortex breakdown are two important phenomena, which determine the effective flow control strategies. The reattachment location on the wing surface moves inboard with increasing angle of attack, and reaches the wing centreline at a particular incidence. Beyond this limiting angle of attack FR, flow reattachment to the wing surface is not possible. The variation of predictions [1] of FR with wing sweep angle is shown in Figure 1. This prediction is only valid for slender wings, and the only experimental observation [2] for a nonslender wing (by visualization of the reattachment line for a sweep angle of =50) is also given in Figure 1. For slender wings, it is seen that the reattachment incidence decreases with increasing wing sweep angle. Therefore, on highly swept wings, reattachment does not occur beyond small angles of attack. Vortex breakdown appears on the wing with increasing angle of attack and crosses the trailing-edge at a particular incidence BD. The variation of this angle of attack
R. King (Ed.): Active Flow Control, NNFM 95, pp. 137151, 2007. springerlink.com Springer-Verlag Berlin Heidelberg 2007

138

I. Gursul et al.

[3] with wing sweep angle is also shown in Figure 1. It is seen that this angle of attack increases with increasing wing sweep angle. On nonslender wings, vortex breakdown appears at very small incidences. As the vortex control concepts become increasingly diverse, new actuators and closed-loop control strategies are being developed. It is useful to consider the flow physics and dominant mechanisms as these determine which flow control methods are effective. The main objective of this paper is to review the vortex control concepts, which mainly depend on the wing sweep.

ANGLE OF ATTACK [deg]

40

30

N W DO N AK W RE B DO X AK TE RE OR B V NO

20
NO RE AT RE TA AT CH TA ME CH NT ME NT

10

0 45

50

55

60

65

70

75

80

85

WING SWEEP [deg]

Fig. 1. Boundaries of vortex breakdown and flow reattachment on the wing surface as a function of sweep angle

2 Slender Delta Wings


2.1 Overview of Flow Physics The flow over a delta wing is characterised by a pair of counter-rotating leading-edge vortices that are formed by the roll-up of vortex sheets. The time-averaged axial velocity is jet-like at low and moderate incidences. The large axial velocities in the vortex core are due to very low pressures, which also generate additional suction and lift force, known as vortex lift, on the delta wings. Vortex lift contribution increases with wing sweep angle [4]. At a sufficiently high angle of attack, the vortices undergo a sudden expansion known as vortex breakdown. The axial flow downstream becomes wake-like with very low velocities. For slender wings (defined as 65 in this paper), vortex breakdown is the dominant flow mechanism that is responsible for decreased lift. It is also the dominant source of unsteadiness that causes wing and fin buffeting [5]. Hence control of vortex breakdown has been the subject of many investigations [6]. There are two important parameters affecting the occurrence and movement of vortex breakdown: swirl level and pressure gradient affecting the vortex core. An increase in the magnitude of either parameter promotes the earlier occurrence of breakdown. Very early experiments [7] demonstrated that vortex breakdown moves upstream over delta wings when the magnitude of either parameter is increased.

Control of Wing Vortices

139

More recently, it was shown [8] that the minimum swirl level required for breakdown decrease with increasing magnitude of adverse pressure gradient. Naturally, flow control methods for the delay of vortex breakdown rely on modification of these two parameters. Active flow control methods generally rely on unsteady excitation in flow control applications. Unsteady forcing has been used for the control of vortex and breakdown in some cases. There are various sources of unsteadiness [5]: shear layer instabilities, vortex wandering, helical mode instability of vortex breakdown, oscillations of breakdown, vortex interaction, and vortex shedding. The frequency spectrum of the unsteady flow phenomena that exist over stationary wings is very wide, which is one of the challenges in numerical simulations of these flows. Vortex breakdown, vortex interactions, and vortex shedding, either alone or in combination, play an important role in wing and fin buffeting, although vortex breakdown is the main source of buffeting over slender wings. Flow control approaches for vortex breakdown are reviewed next. 2.2 Control Surfaces Various control surfaces [6] have been investigated to control the formation, location, strength, and breakdown of the leading-edge vortices: canards, strakes, leading-edge flaps, apex flap, and variable-sweep wings. It is well known that canards can provide substantial delays [9] in vortex breakdown by affecting the external pressure gradient acting on the vortex core. Since all of the vorticity of leading-edge vortices originates from the separation point along the leading-edge, leading-edge flaps are particularly attractive tools that can be used to influence the strength and structure of these vortices. Leading-edge flaps that are deflected downward have been found to reduce drag and improve the lift-to-drag ratio [10]. On the other hand, the flap deflection in the upward direction causes an increase in lift as well as drag. This type of vortex management can be used for landing or aerodynamic manoeuvres [10]. It has been found that the flaps deflected upward generate a stronger vortex lift at low and moderate angles of attack. However, flaps may also induce vortex breakdown [11]. Leading-edge flaps modify the strength and location of the vortices, thereby affecting the parameters that control vortex breakdown. It was shown [11] that breakdown location and its sensitivity strongly depend on incidence and flap angle. For large angles of attack, the variation of breakdown location is not monotonic, and therefore, not suitable for control purposes. A variable leading-edge extension [12] that effectively varies the sweep angle has been used to control leading-edge vortices and breakdown. The advantage of this method is that the variation of breakdown with sweep angle is monotonic, hence suitable for active control purposes. Because most of the vorticity within the vortex core originates from a small region near the apex of the wing, an apex flap can be an effective control surface. It was shown [13] that a drooping apex flap could delay vortex breakdown.

140

I. Gursul et al.

2.3 Blowing and Suction Blowing and suction applied at various locations are commonly used as flow control methods for leading-edge vortices and breakdown. The most widely used versions include: (a) leading-edge suction/blowing, (b) trailing-edge blowing, (c) along-thecore blowing. These are discussed in detail below. Since the vorticity of the leading-edge vortices originates from the separation line along the leading-edge, control of separation characteristics or shear layer can be used to influence the strength and location of the vortices as well as the location of vortex breakdown. Steady blowing [14, 15] and suction [15, 16] at the leading-edge has been employed, but these methods differ in terms of their effect on swirl level. While blowing, in particular tangential blowing, may increase the swirl level, suction reduces the strength and swirl level due to removal of some of the vorticity shed from the leading-edge. Figure 2(a) shows how the location of shear layer and vortex is modified upstream of breakdown when suction is applied. Figure 2(b) shows the axial velocity contours at the trailing-edge, which show the change from wake-like to jet-like velocity as a result of the delay of vortex breakdown. Detailed measurements [16] show that maximum swirl angle in the core and circulation decrease with suction, which causes the vortex breakdown location to move downstream. It is also worth noting that the leading-edge suction technique does not require thick rounded leading edges. Control of vortices can be achieved without the use of the Coanda jet effect.

(a)

(b)

Fig. 2. Variation of (a) rms axial velocity upstream of breakdown location; (b) time-averaged axial velocity at x/c = 1.0

The effect of trailing-edge jets on wing vortices and vortex breakdown has been investigated in several studies [17-20]. Blowing at the trailing-edge also modifies the external pressure gradient and causes delay of vortex breakdown. Favourable effect of a trailing-edge jet [17] could be observed even in the presence of a fin, which produces a strong adverse pressure gradient for a leading-edge vortex. It was shown that fin-induced vortex breakdown can be delayed even for the head-on collision of the leading-edge vortex with the fin [19]. Hence the adverse pressure gradient caused by the presence of the fin could be overcome with a trailing-edge jet. The

Control of Wing Vortices

141

effectiveness of a trailing-edge depends on the wing sweep angle [20]. It appears that it becomes more difficult to delay vortex breakdown with decreasing sweep angle. Along-the-core blowing [21, 22] accelerates the axial flow in the core, and modifies the pressure gradient favourably. Figure 3 shows the effectiveness of various blowing/suction methods from various studies published in the literature. Here the effectiveness is defined as ( xbd/c)/C , where xbd is the change in breakdown location (positive corresponding to delay), c is the chord length and C is the momentum coefficient. Figure 3 shows that along-the-core blowing is the most effective method in terms of delaying vortex breakdown. This can be attributed to the importance of the pressure gradient affecting the vortex core. It is also seen that the effectiveness of blowing and suction is nearly the same. The effectiveness of trailingedge blowing is the lowest among all blowing methods considered.
10
2

Along-the-core blowing

10

Leading-edge suction/blowing

Xbd/c C

100
Trailing-edge blowing

10-1

10

-2

Fig. 3. Effectiveness of various blowing/suction methods from several studies published in the literature

2.4 Unsteady Control There have been various attempts to control vortices and breakdown by using unsteady excitation. These include small and large amplitude oscillations of leadingedge flaps, periodic variations of sweep angle, periodic suction or blowing, and combined use of leading-edge flaps and intermittent trailing-edge blowing, which are summarized in Reference [5]. These studies fall into two categories: (i) highfrequency excitation, St = fc/U = O(1), (where f is the frequency and U is the free stream velocity) and (ii) low-frequency excitation, St = O(0.1). For the high-frequency excitation, Gad-el-Hak and Blackwelder [23] applied periodic perturbations of injection and suction along the leading-edge of a delta wing ( = 60). They found maximum changes in the evolution of the shear layer when the frequency of perturbations (St = 5.5) is the subharmonic of the frequency of KelvinHelmholtz instability. However, no results were reported regarding the structure of the main vortex and the effect on vortex breakdown. Gu et al. [24] applied periodic suction-blowing in the tangential direction along the leading-edge of the wing ( = 75) and reported a delay of vortex breakdown. The most effective period of the alternate suction-blowing corresponded to fc/U = 1.3. For a less slender wing

142

I. Gursul et al.

( = 60), it was shown that oscillatory blowing at the leading edge can enhance the lift at high angles of attack [25], and optimum reduced frequency varied in the range of fc/U = 1 to 2. For small amplitude flap oscillations, the strength of the vortices was larger than that of the quasi-steady case [26], when the excitation frequency was St = 1.2. This range of effective frequencies presumably corresponds to subharmonics of the Kelvin-Helmholtz instability due to vortex pairing. For the low-frequency excitation, there have been reports [5] of increased vortex circulation and delay of vortex breakdown for oscillating flaps, delay of vortex breakdown for harmonic variations of sweep angle for a variable sweep wing, and delay of breakdown for combined flaps and unsteady trailing-edge blowing. When the spectra of unsteady flow phenomena [5] are considered, it is less clear which unsteady phenomena are exploited for the low-frequency excitation. However, it is suggested [27] that the variations in the external pressure gradient generated by unsteady excitation plays a major role. 2.5 Feedback Control The pressure fluctuations induced by the helical mode instability of vortex breakdown can be measured and used as a feedback signal for active control. In such a approach, the rms value of pressure is chosen as the control variable and a feedback control strategy is considered [28]. The monotonic variation of the amplitude of the pressure fluctuations with vortex breakdown location makes the feedback control possible. In identifying a suitable flow controller, several methods were considered, including blowing, suction, and flaps. However, the relationship between the vortex breakdown location and control parameter is unknown or undesirable (i.e., not monotonic) for these methods. A desirable controller should have a monotonic relationship between the control parameter and breakdown location. It was shown [28] that it was feasible to use a variable sweep angle mechanism as a means of controlling the breakdown location by influencing the circulation of the leading-edge vortex. The system was idealized as a first-order system, and integral control was used. The feedback control of breakdown was demonstrated for stationary as well as pitching delta wings.

3 Nonslender Delta Wings


3.1 Overview of Flow Physics Vortical flow over nonslender delta wings ( 55) has recently become a topic of increased interest in the literature. While the flow topology over more slender wings, typically 65, has been extensively studied and is now reasonably well understood, the flow over lower sweep wings has only recently attracted more attention [29]. Vortical flows develop at very low angles of attack, and form close to the wing surface. One of the distinct features of nonslender wings is that reattachment of the separated flow is possible even after breakdown reaches the apex of the wing. However, at large angles of attack in the post-stall region, reattachment

Control of Wing Vortices

143

is not possible, and completely stalled flow occurs on the wing. Active and passive control of reattachment may be beneficial for lift enhancement in the post-stall region. According to Polhamus leading-edge suction analogy [4], the vortex lift contribution becomes a smaller portion of the total lift as the sweep angle decreases. Vortex breakdown occurs over the wing even at small incidences, and there is no obvious correlation between the onset of vortex breakdown over nonslender wings and the change of the lift coefficient. Hence, vortex breakdown is not a limiting phenomenon as far as the lift force is concerned for nonslender wings. On the contrary, flow reattachment is key to any flow control strategy as suggested in Figure 1. 3.2 Passive Control with Wing Flexibility Passive lift enhancement for flexible delta wings has been demonstrated as a potential method for the control of vortex-dominated wing flows [30]. Force measurements over a range of nonslender delta wings (with sweep angles = 40 to 55) have demonstrated the ability of a flexible wing to enhance lift and delay stall compared with a rigid wing of similar geometry. An example for = 40 is shown in Figure 4a. This recently discovered phenomenon appears to be a feature of nonslender wings. Flow visualization, PIV and LDV measurements show that flow reattachment takes place on the flexible wings in the post-stall region of the rigid wings. The lift increase in the post-stall region is accompanied with large self-excited vibrations of the wings as shown in Figure 4b. The dominant frequency of the vibrations of various nonslender delta wings is St = O(1). These vibrations promote reattachment of the shear layer, which results in the lift enhancement. Various measurements including wing-tip accelerations, rms rolling moment, and hot-wire measurements confirm that the dominant mode of vibrations occur in the second antisymmetric structural mode in the lift enhancement region. The self-excited vibrations are not observed for a halfwing model, hence passive flow control for a flexible wing occurs only in the anti-symmetric mode.
1.5

(a)

1.4 1.3 1.2 1.1 1 0.9

(b)

0.2 0.18 0.16 0.14 0.12

Mean deflection Peak-to-peak amplitude

CL

/s
Flexible Rigid

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 20

0.1 0.08 0.06 0.04 0.02

25

30

35

40

10

15

20

25

30

35

40

45

Fig. 4. (a) Variation of lift coefficient as a function of angle of attack for rigid and flexible wings, = 40; (b) Variation of mean and amplitude of fluctuating wing-tip deflection with incidence, =40

144

I. Gursul et al.

Spectral analysis of velocity fluctuations [31] along the shear layer showed large sharp peaks, corresponding to the wing vibrations. There are also broad dominant peaks in the spectra of velocity fluctuations in the range of St = 1 to 5 for the poststall incidences, and these correspond to the shear layer instabilities. The center frequency of these peaks decreases with streamwise distance as the shear layer vortices shed conically. There is also a decrease in the spanwise direction due to the vortex pairing process. The frequency of the structural vibrations is in the same range as these natural frequencies [31]. 3.3 Active Control of Reattachment Rigid delta wings undergoing small amplitude oscillations [32] in the post-stall region exhibit many similarities to flexible wings, including reattachment in the post-stall region (see Figure 5). For simple delta wings and cropped delta wings [31] with = 50, 40, and 30, an optimum frequency around St = 1 was identified for which the reattachment is observed. Note that these dramatic changes are observed with unsteady forcing in the post-stall region, whereas there is little effect in the pre-stall region. Hence active control methods in the form of leading-edge oscillations or blowing can be effective. An important parameter is the wing sweep angle. The effect of excitation on a swept wing is similar to the response of the flow over a backward-facing step [33] to the periodic excitation. However, for zero sweep angle, formation of a closed separation bubble at high angles of attack in the post-stall region is not possible. It seems that moderate sweep angles (around 50) help the formation of semi-open separation bubbles, hence the wing sweep is beneficial in flow reattachment. However, there is a lower limit of sweep angle below which the beneficial effect of wing sweep will diminish. This lower limit of sweep angle is around = 20. Symmetric perturbations in the form of small amplitude pitching oscillations (1 amplitude) were studied for = 50 simple delta wing. The results show that symmetric perturbations also promote reattachment and vortex re-formation. Hence, St=0
(a)

St=1.0

(b)

Fig. 5. Effect of shear layer excitation for =25 and =50; (a) time-averaged fluorescent dye visualization, (b) instantaneous visualization for one side only

Control of Wing Vortices

145

for active control purposes, both symmetric and anti-symmetric excitations are effective. However, passive control for a flexible wing occurs only in the antisymmetric mode. 3.4 Vortex Re-formation Within the reattachment region, axial flow may develop, resulting in re-formation of the leading-edge vortices. Figure 6 shows an example of vortex re-formation [32], which occurs when the amplitude of forcing is sufficiently large and the frequency of excitation is near an optimum value. The mean breakdown location becomes a maximum at an optimum frequency as shown in Figure 7.

fc/U = 0

fc/U = 1.8

Fig. 6. Flow visualization for stationary and oscillating wings in water tunnel, = 50, =25

0.3

= 25, = 5 = 25, = 1

0.2

xbd/c
0.1 0 0

10

fc/U

Fig. 7. Variation of time-averaged breakdown location as a function of dimensionless frequency for two values of oscillation amplitude, = 50, =25

The time-averaged vorticity flux increases due to the oscillating leading edge, which leads to increased circulation. Although the leading edge vortices become stronger due to the leading edge motion, vortex breakdown is delayed for the oscillating wing compared to the stationary wing for which breakdown is at the apex. This appears to be in contrast to the well-known studies of vortex breakdown, which indicate that increased strength of vortices should cause premature, rather than

146

I. Gursul et al.

delayed, breakdown. This result suggests that streamwise pressure gradient might be modified favourably due to the wing motion.

4 High Aspect Ratio Wings


4.1 Vortex Control Concepts Control of formation and development of tip vortices in both near-wake and far-wake has potential benefits in a variety of aerodynamic problems. Tip vortices form in a similar way to the leading-edge vortices over low aspect ratio wings and the roll-up process becomes complete within a few chord-lengths downstream of the trailingedge. There appear to be three main applications of flow control approaches with regard to the tip vortices: 1) reduction of induced drag, 2) attenuation of vortex wake hazard on following aircraft, 3) reduction of helicopter noise due to the blade-vortex interaction. These will be briefly reviewed with regard to vortex control concepts. For reduction of induced drag, various methods were considered [34], including planform shape and span, tip shape, winglets, fences and tip sails. Wing tip devices are used to redistribute the vorticity near the wing tip. Similarly, any flow control method that can displace the tip vortices in the outboard direction can be used to reduce the drag as it is inversely proportional to the square of the spanwise distance between the vortices. In order to attenuate wake vortex hazard on following aircraft, the core size of the vortices can be increased by turbulence injection into the core. Various wing tip devices were tested, which were found to be not much effective in the far-wake. Those that seemed to decrease the vortex hazard (such as decelerating chutes and splines placed behind the trailing-edge) had an unacceptable drag penalty [35]. It was noted [36] that passive devices in the form of turbulence generators could increase the size of the vortex core and also promote cooperative instabilities as an additional benefit. A second and more promising method is to rely on the longwave instabilities occurring in a system of several vortices [37]. Compared to the alleviation of a single vortex by turbulent diffusion, excitation of the long-wave cooperative instabilities of multiple vortices has the potential for much faster destruction of vortex wakes. Hence deliberate creation of multiple vortices and excitation by oscillating surfaces or oscillatory blowing may be a promising strategy. One of the sources of helicopter noise is the interaction of rotor blades with the vortices shed from preceding blades. Noise and vibration caused by this interaction can be reduced by i) increasing the distance between the rotor blade and tip vortex, and (ii) by increasing the size of the vortex core and decreasing the maximum tangential velocity. 4.2 Tip Blowing Most of the intended modifications to the tip region (such as wing tip devices), to the vortex location, strength, and structure can be achieved without the use of passive devices. Active flow control using wing tip blowing can achieve multiple tasks

Control of Wing Vortices

147

during different flight regimes. It was shown that the strength, location, core structure, and number of vortices can be effectively manipulated by tip blowing [38]. The effect of continuous blowing using high-aspect ratio jets is very sensitive to the blowing direction. The location and strength of the vortices generated by the jet and their interaction with the tip vortex lead to different flow configurations. Blowing in the upward direction produces additional co-rotating vortices in the nearwake as shown in Figure 8. These strong vortices forming on the wing surface can be used to increase the lift in certain applications such as hovering rotorcraft. Figure 9 shows that a diffused vortex is obtained with spanwise blowing near the pressure surface. Also, downward blowing produces diffused vortices. For these cases, blowing appears to add a substantial amount of turbulence in the vortex core, resulting in diffused trailing vortices. For downward blowing, it was possible to displace the vortices in the outboard direction, which should result in an induced drag

Fig. 8. Vorticity distribution for upward blowing

Fig. 9. Vorticity distribution for spanwise blowing near the pressure surface

148

I. Gursul et al.
0.3

(b)

1.2

0.2

1
0.1

U /U
0.8

z/c

synthetic jet - C =0.0004 synthetic jet - C =0.0008 continuous jet - C =0.0004


-0.1 -0.1 0

y/c

0.1

0.2

0.3

0.6

10-1

fc/U

100

101

Fig. 10. (a) Velocity and vorticity for the synthetic jet (with no freestream); (b) variation of maximum tangential velocity at x/c=1.0 as a function of dimensionless frequency

reduction. Spanwise blowing also results in a lift augmentation [39] as the downwash decreases with blowing. Drag reduction and lift augmentation are essentially inviscid phenomena due to an effective increase in span and aspect ratio of the wing. Synthetic jets can be useful for wing tip blowing as the air supply problem is avoided. Figure 10 shows that a synthetic jet is beneficial in terms of diffusing the trailing vortex [40]. Synthetic jets produce large turbulence and much larger vortex wandering. For the blowing coefficients used, the effect of excitation frequency was minor. While there are substantial changes in the circulation of the trailing vortex for continuous jets, the circulation remains virtually the same for synthetic jets. The intermittent and diffused jet vortices are weak and do not appear to affect the total circulation.

5 Conclusions
For slender delta wings, control of vortex breakdown has been the primary goal of many investigations. Delay of vortex breakdown is possible with the modifications to the swirl level and pressure gradient. The use of control surfaces such as leadingedge flaps makes it possible to control the location, strength, and structure of the vortices. Blowing and suction at the leading-edge, trailing-edge, or along the core have differences in terms of their effects on swirl level and pressure gradient affecting the vortex core. Along-the-core blowing is the most effective method for delaying vortex breakdown. Active flow control using high-frequency excitation targets the Kelvin-Helmholtz instability of the separated shear layers. In contrast, low-frequency excitation appears to modify the external pressure gradient only. For nonslender delta wings, flow reattachment is the most important aspect for flow control methods. Passive lift enhancement on flexible wings is due to the selfexcited wing vibrations, which promote flow reattachment in the post-stall region. The frequency of the wing vibrations (St = O(1)) is in the same range as the natural frequencies of the shear layer instabilities. Rigid delta wings undergoing small amplitude oscillations in the post-stall region exhibit many similarities to flexible wings, including reattachment and vortex re-formation. Moderate sweep angles help the formation of semi-open separation bubbles, hence the wing sweep is beneficial.

Control of Wing Vortices

149

For high aspect ratio wings, vortex control concepts are diverse, ranging from drag reduction to attenuation of wake hazard and noise. Modifications to the vortex location, strength, and structure are common ideas in these applications. Active flow control using wing tip blowing is shown to have potential for various objectives. Depending on the blowing configuration and direction, more diffused or stronger vortices, and also, multiple vortices, that move inboard or outboard can be generated. The deliberate creation of multiple vortices and excitation of vortex instabilities may be a promising strategy for effective control in the far-wake.

Acknowledgements
The authors acknowledge the financial support of the Air Force Office of Scientific Research (AFOSR), Engineering and Physical Sciences Research Council (EPSRC) and the Ministry of Defence in the UK.

References
[1] Mangler, K.W. and Smith, J.H.B., A Theory of the Flow Past a Slender Delta Wing with Leading Edge Separation, Proceedings of Royal Society, A, vol. 251, 1959, pp. 200-217. [2] Taylor, G. and Gursul, I., Buffeting Flows over a Low Sweep Delta Wing, AIAA Journal, vol. 42, no. 9, September 2004, pp. 1737-1745. [3] Erickson, G.E., Water-Tunnel Studies of Leading-Edge Vortices, Journal of Aircraft, vol. 19, No. 6, June 1982, pp. 442-448. [4] Polhamus, E. C., Predictions of vortex lift characteristics by a leading edge suction analogy, Journal of Aircraft, Vol. 8, No. 4, pp. 193-199, 1971. [5] Gursul, I., Review of Unsteady Vortex Flows over Slender Delta Wings, Journal of Aircraft, vol. 42, no. 2, March-April 2005, pp. 299-319. [6] Mitchell, A.M. and Delery, J., Research into Vortex Breakdown Control, Progress in Aerospace Sciences, vol. 37, 2001, pp. 385-418. [7] Lambourne, N.C. and Bryer, D.W., The Bursting of Leading Edge Vortices: Some Observation and Discussion of the Phenomenon, Aeronautical Research Council, R&M 3282, 1962. [8] Delery, J., Horowitz, E., Leuchter, O., and Solignac, J.L., Etudes Fondamentales Sur Les Ecoulements Tourbillonnaires, La Recherche Aerospatiale, no. 2, 1984, pp. 81-104. [9] Myose, R.Y., Hayashibara, S., Yeong, P.C. and Miller, L.S., Effects of Canards on Delta Wing Vortex Breakdown During Dynamic Pitching, Journal of Aircraft, vol. 34, no. 2, March-April 1997, pp. 168-173. [10] Lamar, J.E. and Campbell, J.F., Vortex Flaps Advanced Control Devices for Supercruise Fighters, Aerospace America, January 1984, pp. 95-99. [11] Deng, Q. and Gursul, I., Effect of Leading-Edge Flaps on Vortices and Vortex Breakdown, Journal of Aircraft, vol. 33, no. 6, November-December 1996, pp. 10791086. [12] Yang, H. and Gursul, I., Vortex Breakdown over Unsteady Delta Wings and its Control, AIAA Journal, vol. 35, no. 3, 1997, pp. 571-574. [13] Klute, S.M., Rediniotis, O.K., and Telionis, D.P., Flow Control over a Maneuvering Delta Wing at High Angle of Attack, AIAA Journal, vol. 34, no. 4, 1996, pp. 662-668.

150

I. Gursul et al.

[14] Wood, N.J., Roberts, L. and Celik, Z., Control of Asymmetric Vortical Flows over Delta Wings at High Angles of Attack, Journal of Aircraft, vol. 27, no. 5, May 1990, pp. 429435. [15] Gu, W., Robinson, O. and Rockwell, D., Control of Vortices on a Delta Wing by Leading-Edge Injection, AIAA Journal, vol. 31, no. 7, July 1993, pp. 1177-1186. [16] McCormick, S. and Gursul, I., Effect of Shear Layer Control on Leading Edge Vortices, Journal of Aircraft, vol. 33, no. 6, November-December 1996, pp. 1087-1093. [17] Helin, H.E. and Watry, C.W., Effects of Trailing-Edge Jet Entrainment on Delta Wing Vortices, AIAA Journal, vol. 32, no. 4, 1994, pp. 802-804. [18] Shih, C. and Ding, Z., Trailing-Edge Jet Control of Leading-Edge Vortices of a Delta Wing, AIAA Journal, vol 34, no(7), 1996, 1447-1457. [19] Phillips, S., Lambert, C., and Gursul, I., Effect of a Trailing-Edge Jet on Fin Buffeting, Journal of Aircraft, vol. 40, no. 3, 2003, pp. 590-599. [20] Wang, Z. and Gursul, I., Effects of Jet/Vortex Interaction on Delta Wing Aerodynamics, 1st International Conference on Innovation and Integration in Aerospace Sciences, 4-5 August 2005, Queens University Belfast, UK. [21] Guillot, S., Gutmark, E.J., and Garrison, T.J., Delay of Vortex Breakdown over a Delta Wing via Near-Core Blowing, AIAA 98-0315, 36th Aerospace Sciences Meeting and Exhibit, January 12-15, 1998, Reno, NV. [22] Mitchell, A.M., Barberis, D., Molton, P., and Delery, J., Oscillation of Vortex Breakdown Location and Blowing Control of Time-Averaged Location, AIAA Journal, vol. 38, no. 5, May 2000, pp. 793-803. [23] Gad-el-Hak, M. and Blackwelder, R.F., Control of the Discrete Vortices from a Delta Wing, AIAA Journal, vol. 25, no. 8, 1987, pp. 1042-1049. [24] Gu, W., Robinson, O. and Rockwell, D., Control of Vortices on a Delta Wing by Leading-Edge Injection, AIAA Journal, vol. 31, no. 7, July 1993, pp. 1177-1186. [25] Margalit, S., Greenblatt, D., Seifert, A. and Wygnanski, I., Delta Wing Stall and Roll Control Using Segmented Piezoelectric Fluidic Actuators, Journal of Aircraft, vol. 42, no. 3, 2005, pp. 698-709. [26] Deng, Q. and Gursul, I., Effect of Oscillating Leading-Edge Flaps on Vortices over a Delta Wing, AIAA 97-1972, 28th AIAA Fluid Dynamics Conference, June 29 July 2, 1997, Snowmass Village, CO. [27] Yang, H. and Gursul, I., Vortex Breakdown over Unsteady Delta Wings and Its Control, AIAA Journal, vol. 35, no. 3, 1997, pp. 571-574. [28] Gursul, I., Srinivas, S. and Batta, G., Active Control of Vortex Breakdown over a Delta Wing, AIAA Journal, vol. 33, no. 9, 1995, pp. 1743-1745. [29] Gursul, I., Gordnier, R., and Visbal, M., Unsteady Aerodynamics of Nonslender Delta Wings, Progress in Aerospace Sciences, vol. 41, 2005, pp. 515-557. [30] Taylor, G., Kroker, A. and Gursul, I., Passive Flow Control over Flexible Nonslender Delta Wings, AIAA-2005-0865,43rd Aerospace Sciences Meeting and Exhibit Conference,10-13 January 2005,Reno,NV. [31] Gursul, I., Vardaki, E. and Wang, Z., Active and Passive Control of Reattachment on Various Low-Sweep Wings, AIAA-2006-506, 44th AIAA Aerospace Sciences Meeting and Exhibit, 9-12 January 2006, Reno, NV. [32] Vardaki, E., Gursul, I. and Taylor, G., Physical Mechanisms of Lift Enhancement for Flexible Delta Wings, AIAA-2005-0867, 43rd Aerospace Sciences Meeting and Exhibit, 10-13 January 2005, Reno, NV.

Control of Wing Vortices

151

[33] Roos, F.W. and Kegelman, J.T., Control of Coherent Structures in Reattaching Laminar and Turbulent Shear Layers, AIAA Journal, vol. 24, no. 12, December 1986, pp. 19561963. [34] Kroo, I., Drag Due to Lift: Concepts for Prediction and Reduction, Annual Review of Fluid Mechanics, vol. 33, 2001, pp. 587-617. [35] Spalart, P.R., Airplane Trailing Vortices, Annual Review of Fluid Mechanics, vol. 30, 1998, pp. 107-138. [36] Coustols, E., Stumpf, E., Jacquin, L., Moens, F., Vollmers, H., Gerz, T., Minimised Wake: a Collaborative Research Programme on Aircraft Wake Vortices, AIAA 2003-0938, 41st Aerospace Sciences Meeting and Exhibit, 6-9 January 2003, Reno, NV. [37] Jacquin, L., Fabre, D., Sipp, D., Theofilis, V., and Vollmers, H., Instability and Unsteadiness of Aircraft Wake Vortices, Aerospace Science and Technology, vol. 7, 2003, pp. 577-593. [38] Margaris, P., Gursul, I., Effect of Steady Blowing on Wing Tip Flowfield, AIAA 20042619. 2nd Flow Control Conference, Portland, Oregon, USA. June-July 2004. [39] Tavella, D. A., Wood, N J., Lee, C. S., Roberts, L., Lift Modulation with Lateral WingTip Blowing, Journal of Aircraft, Vol. 25, No. 4, 1988, pp 311-316. [40] Margaris, P., Gursul, I., Wing Tip Vortex Control Using Synthetic Jets, CEAS/KATnet Conference on Key Aerodynamic Technologies, 20-22 June 2005, Bremen, Germany.

Das könnte Ihnen auch gefallen