Vous êtes sur la page 1sur 9

38th Fluid Dynamics Conference and Exhibit<BR> 23 - 26 June 2008, Seattle, Washington

AIAA 2008-3834

Laminar Flow Control Flight Tests for Swept Wings: Strategies for LFC
William S. Saric 1 , Andrew L. Carpenter. 2 , and Helen L. Reed 3 Texas A&M University, College Station, Texas 77843-3141

A brief review of Laminar Flow Control (LFC) techniques are given and a strategy for achieving laminarization for transonic transports is discussed. A review of some flight test results on swept-wing transition are presented. It is shown that polished leading edges can create large regions of laminar flow because the flight environment is relatively turbulence free and the surface finish reduces the initial amplitude of the stationary crossflow vortex.

I. Laminar Flow Control (LFC)


Laminar-turbulent transition occurs as the result of the growth of unstable disturbances within the boundary layer. The principle behind LFC is to keep the growth of these disturbances within acceptable limits so that 3-D and nonlinear effects do not cause breakdown to turbulence. With this philosophy, one only works with linear disturbances whose behavior can be reliably calculated. Thus, the difficulties with transition prediction do not directly arise. Bushnell & Tuttle (1979) and Bushnell (1992) are good reviews of the subject. A. Streamwise Instabilities: Tollmien-Schlichting (T-S) Waves The manner in which LFC works can be described using the following example of Reshotko (1984, 1985). It is well known that the velocity-profile curvature term in the Orr-Sommerfeld equation, d 2U dy 2 , is an important driver of the stability behavior. In fact, it is more important than immeasurable changes in the mean velocity itself. The boundary-layer flow can be made more stable by making the curvature term more negative near the wall. This results in a fuller profile that approaches a turbulent profile. The boundary-layer momentum equation, with = (T), can be evaluated near the wall, as shown in Eq. (1), and used to illustrate the stabilizing effects of different LFC techniques.

( V0 )

U P d T U 2U + = 2 as y 0 y x dT y y y

(1)

Here, is the density, V0 is the wall-normal velocity evaluated at the wall (+ for blowing, - for suction), U is the streamwise velocity component, P is the pressure field, is the dynamic viscosity, T is the temperature, x is the streamwise coordinate, and y is the wall-normal coordinate. An evaluation of (1) shows that wall suction (V0 < 0 ) , an accelerating pressure gradient ( P x < 0 ) , wall cooling in air ( d dT > 0, T y > 0 ) , and wall heating in water ( d dT < 0, T y < 0 ) , all tend to stabilize the boundary layer by making the profile curvature term more negative near the wall. The case for pressure gradient: It should be pointed out that these are very sensitive mechanisms and that even weak suction or weak pressure gradients produce strong effects. For example, a Falkner-Skan pressure gradient of = +0.1 (which has only a 6.6% change in the shape factor H = from Blasius), increases the minimum critical x-Reynolds number for stability by a factor of 9. The case for suction: At the same time, average suction velocity ratios of V0 U = 103 to 104 , which are not unusual for LFC applications, can, for example, reduce relative amplitude growth from e 26 to e5 at F = 10 106

1 2

Distinguished Professor, Aerospace Engineering, MS-3141, Fellow AIAA. Graduate Student, Aerospace Engineering, MS-3141, Student Member AIAA. 3 Professor and Head, Aerospace Engineering, MS-3141, Fellow AIAA. 1 American Institute of Aeronautics and Astronautics
092407
Copyright 2008 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

(Saric & Nayfeh 1977). That the idea works is evidenced by the fact that the X-21 aircraft achieved laminar flow at Reynolds numbers of 27 x 106 with a 20% decrease in overall drag (Pfenninger 1977). The Orr-Sommerfeld Equation (OSE) is solved with the appropriate basic state used as input (Reed & Nayfeh 1986). Justification for using the usual homogeneous boundary conditions on a disturbance velocity over the porous or perforated sections follows the work of Lekoudis (1978). The experiments (Reynolds & Saric 1986) justify all aspects of the theory (Reed & Nayfeh 1986). The case for heating and cooling: Examples of the heating and cooling applications in LFC systems are reviewed by Reshotko (1978, 1979, 1984, 1985). Successful LFC by heating in water has been demonstrated (Lowell & Reshotko 1974; Strazisar et al 1977; Barker & Gile 1981) but the practicality of using the technique in other than particulate-free, fresh water is in question. ). Reed & Balakumar (1990) and Balakumar & Reed (1991) calculate stability for cooling in air and show that the cooling stabilizes the usual supersonic T-S mode but destabilizes the acoustic mode (second mode: Mack 1984). Reshotko (1979) proposed a scheme for cooling in air but transport LFC applications have not yet left the laboratory. A strong case cannot be made for heating or cooling LFC. Hybrid Laminar Flow Control: Present designs for energy-efficient airfoils have LFC systems with a porous region near the leading edge. Generally, suction is applied near the leading edge of a swept wing in order to control leading-edge contamination and crossflow instabilities. Appropriate shaping of the pressure distribution stabilizes mid-chord instabilities. This arrangement is called a hybrid LFC system (HLFC) in that it combines steady-suction LFC with passive LFC (pressure gradient). A successful example of HLFC was conducted by Boeing (1993). B. Active control of transition The idea of transition control through active feedback systems is an area that has received considerable attention (Liepmann & Nosenchuck 1982; Thomas 1983; Pupator & Saric 1989). The technique consists of first sensing the amplitude and phase of an unstable disturbance and then introducing an appropriate out-of-phase disturbance that cancels the original disturbance. In spite of some early success, this method is no panacea for the transition problem. Besides the technical problems of the implementation of such a system on an aircraft, the issue of three-dimensional wave cancellation must be addressed. Although Pupator & Saric (1989) successfully canceled broadband random 2-D waves with an active feedback system, Thomas (1983) showed that, when the 2-D wave is canceled, all of the features of the 3-D disturbances remain to cause transition at yet another location. Some clear advantage over passive systems has yet to be demonstrated for any active feedback technique. C. Three-dimensional boundary layers The crossflow instability occurs in regions of pressure gradient on swept surfaces. In the inviscid region outside the boundary layer, the combined influences of sweep and pressure gradient produce curved streamlines at the boundary-layer edge. Inside the boundary layer, the streamwise velocity is reduced, but the pressure gradient is unchanged. Thus, the same balance between centripetal acceleration and pressure gradient that is present in the inviscid layer, does not exist does not exist in the boundary layer. This imbalance results in a secondary flow in the boundary layer, called crossflow, that is perpendicular to the direction of the inviscid streamline. The basic tutorial and review on the subject is given by Saric et al (2003). When considering the control of instabilities, the crossflow is the most difficult. Whereas an accelerating Cp is stabilizing to T-S waves, it destabilizes crossflow. Moreover, unlike T-S instabilities, the crossflow problem exhibits amplified disturbances that are stationary as well as traveling. Even though both types of waves are present in typical swept-wing, transition is usually caused by either one, but not both, of these waves. Although linear theory predicts that the traveling disturbances have higher growth rates, transition in many experiments is induced by stationary waves. Whether stationary or traveling waves dominate is related to the receptivity process. Stationary waves are more important in low-turbulence environments characteristic of flight, whereas traveling waves dominate in high-turbulence environments (Bippes 1997; Deyhle & Bippes 1996). Stationary crossflow waves often show evidence of strong nonlinear effects (Dagenhart et al, 1989 1990; Bippes & Nitschke-Kowsky 1990; Bippes et al 1991; Deyhle et al 1993; Radeztsky et al 1994; Reibert et al 1996; Saric et al 1998). Because the wave fronts are fixed with respect to the model and are nearly aligned with the potential-flow direction (i.e., the wavenumber vector is nearly perpendicular to the local inviscid streamline), the weak motion of the wave convects O(1) streamwise momentum producing a strong distortion in the streamwise boundary-layer profile. This integrated effect and the resulting local distortion of the mean boundary layer lead to the modification of the basic state and the early development of nonlinear effects. As the distortions grow, the boundary layer develops an alternating pattern of accelerated, decelerated, and doubly inflected profiles. The inflected profiles are inviscidly unstable and, as such, are subject to a high-frequency secondary instability (Kohama et al 1991; Malik et 2 American Institute of Aeronautics and Astronautics
092407

al 1994; Wassermann & Kloker 2002; White & Saric 2005). This secondary instability is highly amplified and leads to rapid local breakdown. Because transition develops locally, the transition front is nonuniform in span and characterized by a saw-tooth pattern of turbulent wedges. Two important observations concerning the distributed roughness results byReibert et al (1996) are (a) unstable waves occur only at integer multiples of the primary disturbance wave number and (b) no subharmonic disturbances are destabilized. Spacing the roughness elements with wave number k = 2/ excites harmonic disturbances with spanwise wave numbers of 2k, 3k, , nk (corresponding to /2, /3, , /n) but does not produce any unstable waves with intermediate wavelengths or with wavelengths greater than . Following this lead, Saric et al (1998) investigated the effects of periodic distributed roughness whose primary disturbance wave number did not contain a harmonic at = 12 mm (the most unstable wavelength according to linear theory). By changing the forced fundamental disturbance wavelength (i.e., the roughness spacing) to 18 mm, the velocity contours clearly showed the presence of the 18-mm, 9-mm, and 6-mm wavelengths. However, the linearly most unstable disturbance (12 mm) was completely suppressed. Moreover (and consistent with all previous results), no subharmonic disturbances were observed, which shows that an appropriately designed roughness configuration can, in fact, inhibit the growth of the (naturally occurring) most unstable disturbance. When the disturbance wavelength was forced at 8 mm, the growth of all disturbances of greater wavelength was suppressed. The most remarkable result obtained from the subcritical roughness spacing is the dramatic effect on transition location: In the absence of artificial roughness, transition occurs at approximately 55% chord. Adding roughness with a spanwise spacing equal to the wavelength of the linearly most unstable wave moves transition forward to 47% chord. However, subcritical forcing at 8-mm spanwise spacing actually delays transition beyond the pressure minimum and well beyond 80% chord (the actual location was beyond view). Subsequent to the experiments, the NPSE results (Haynes & Reed 2000) confirmed this effect. In a DNS solution, Wassermann & Kloker (2002) have shown the same stabilization due to subcritical forcing. Using the same independent approach in terms of the calculation of the basic state, they demonstrated the stabilization due to subcritical roughness.

II. Strategies for Laminar Flow Control


Because the distributed roughness technique has not been demonstrated at chord Reynolds numbers that are characteristic of transport LFC systems, there has been some heavy weather generated regarding the range of DRE applicability. We should like to clear the air on this matter by listing a possible stepwise strategy for LFC. This list is arranged beginning with the highest Technical Readiness Level (TRL) to the lowest TRL. A. Use of a suction system If one had a real and immediate requirement for LFC on a transonic transport, we should recommend first the use of weak wall suction. Pfenninger made it work on the X-21. In one of the recent flight examples, Boeing (1993) successfully demonstrated suction HLFC on the 757 up to a Rex = 16 x 106. Everyone agrees that suction works for the attachment line, T-S, and crossflow. It is simple and has a strong fundamental basis. Wagner et al (1988) and Maddalon et al (1990) made the case for suction LFC working in an operational environment. B. Two-dimensional wing If one then said that the cost, reliability, structural, and certification issues made suction impractical, we should recommend as the second option, reducing the wing sweep to 10 - 12 and eliminating crossflow instabilities. Then use an accelerating pressure gradient to stabilize T-S. Determine the requirements for efficient pressure recovery and place the pressure minimum as far aft as possible. Pressure gradient stabilization of T-S waves is a reliable passive technique that has been proven to work. Richard Tracy and his colleagues at Reno Air and Aerion have been making this case for years (Matisheck 2007; Sturdza 2007). C. Polished Leading Edge If one then said that the wave-drag penalties or pressure-recovery problems make a 2-D wing impractical, we should then recommend as the third option, having a polished leading edge. The stationary crossflow wave draws its initial amplitude from surface roughness. Saric et al (2008) have demonstrated that with a leading edge polished to 0.3 m rms, laminar flow can be achieved on a 30 swept wing with N-factors up to 14 or so (see a summary below). Thus, laminar flow can be obtained to at least the pressure minimum. One can use a leading-edge Krger (Boeing, 1993) to protect the polished surface during landing and take off. If a material harder than aluminum is plated on the leading edge, the surface finish can be improved and higher N-factors for transition can be achieved. The conjecture is that one would have no difficulty achieving natural laminar flow with a nickel-plated leading edge. 3 American Institute of Aeronautics and Astronautics
092407

D. Periodic Distributed Roughness Elements (DREs) If one then said that the operational requirements prevented a highly-polished leading edge, only then would we suggest the use of periodic DRE technology because there is nothing left on the table to try. The TRL for periodic DRE is not high and significant research remains to be done. The technique has been demonstrated in flight at Rec = 8 x 106. It is suggested that additional flight tests be conducted at transport unit Reynolds numbers (M = 0.8, H = 40 kft) but with a modest Rec in the range of 15 20 x 106. Wind-tunnel tests are not recommended because of the higher turbulence levels and a greater sensitivity to Rek.

III. Flight Test Results


The original wind-tunnel experiments and computations were done in a modest chord-Reynolds-number range (2.2 to 3.5 million) and the goal has been to extend this to higher chord Reynolds numbers more typical of flight systems. Because of the sensitivity of the crossflow instability to freestream turbulence, it appears to be difficult (if not impossible) to do laminar crossflow experiments at higher Reynolds numbers (>8 million) in wind tunnels because of turbulence. Flight tests can be very difficult since one does not have the collection of instrumentation available to a wind tunnel. However, if one follows the guidelines of Transition Study Group for transition research in flight and use the care outlined by Saric (1990), there is a chance for success. The influence of freestream disturbances must be resolved and an important step is to do careful stability and transition experiments in flight where the disturbance levels are indeed low. These experiments should form the base state for the influence of roughness. A well-known and very successful flight program was conducted by Dougherty & Fisher (1980). Since the identical model was taken to every supersonic facility, this work actually provided a means to evaluate flow quality in high-speed tunnels. Since then, the achievements have been meager for a variety of reasons not the least of which is the cost of doing flight experiments. A. Objectives The objective was to investigate, in a low-disturbance, flight-test environment, the periodic DRE technology on a subsonic swept-wing test article. The test article was designed to be consistent with a SensorCraft-type wing section (30 leading-edge sweep). The goals are to quantify the effectiveness of periodic DRE in increasing the extent of laminar flow (i.e. transition location in chordwise direction) on the suction and/or pressure sides beyond the baseline (no-control) case; investigate the robustness and utility of DRE in maintaining laminar flow over the SensorCraft flight envelope i.e., variations in test-article angle-of-attack (AoA) over chord Reynolds numbers, Rec = 7.5 x 106; gain insight into conducting boundary-layer transition control experiments in a flight environment versus a wind-tunnel environment; and obtain a database that provides additional insight into boundary-layer stability and transition and for validation of prediction tools. The AOA was nominally set at 0 but was adjusted to as much as 2 using sideslip. The program planning objectives were: (1) Measure the freestream disturbance environment and establish that the flight test has an acceptable disturbance environment within which one can conduct boundary-layer stability and transition measurements; (2) Develop a map of breakdown due to isolated roughness as a function of Rek and roughness location (Rex); (3) Develop the laminarization technology with periodic DRE and determine the sensitivity to roughness at higher Reynolds numbers; (4) Determine how the low-disturbance environment of flight can validate (or invalidate) wind-tunnel experiments; (5) Complement the experiments with stability computations; (6) Provide program guidelines for laminarization and long-range flight. All six objectives were met. B. Test Results The primary objective for Phase I testing was to determine whether the in-flight turbulence intensities were low enough to proceed with the swept-wing experiment. A value less than 0.08% for u/U was expected. Experimental results show that the nominal value is between 0.05%U and 0.06%U. Basically the target conditions for achieving 60% laminar flow where a chord Reynolds number of Rec =7.5 x 106, at model angle of attack of AoA = -4, and a swept angle of = 30. The model (see Figure 1).was fabricated at Tri-Models in Huntington Beach, California and was flown on a Cessna O-2 as an external store.

4 American Institute of Aeronautics and Astronautics


092407

Figure 1. The swept-wing model hung on O-2. A black powder-coat finish was used to enhance the IR image. The IR camera was mounted in the cabin. C. Initial results with a polished leading edge The swept-wing model was designed with an accelerated flow to 70% chord. The intent was to make the boundary layer sub-critical to T-S waves but rather unstable to crossflow instabilities. One of the principal result is that we achieved 80% laminar flow with a polished leading edge at Rec = 8.1 x 106, AoA = -4 and = 30. This corresponds to linear stability N-factors of well over 14. Background roughness was 0.3 m rms with 2.2 m avg peak-to-peak. The linear stability N-factor is the log of the unstable disturbance amplitude ratio as shown in Eqn. (1).
N = ln A A0

(1)

Where A0 is the initial amplitude at the first neutral point and A is the amplitude at transition. Thus an e14 growth is an amplitude ratio of 106. The IR Thermography for this case is shown in Figure 2.

Figure 2. IR image at 170 KTAS, Rec = 8.0 x 106, AoA = -4, = 30; 3500 ft MSL, Polished LE, No DRE, peak to peak roughness = 4.3 m; rms roughness = 0.33 m, N-factor > 14 at mid-span, x/c)tr = 80% 5 American Institute of Aeronautics and Astronautics
092407

The colder area denoted by the dark orange color indicates laminar flow while the lighter area denotes turbulent flow. These conclusions were confirmed by placing large roughness elements on the model and tripping the boundary layer. The white marks at the bottom and top of the model are pieces of aluminum tape denoting 40%, 60%, and 80% chord respectively. The light orange color near the top of the model is due to the cabin IR reflection. The diagonal line across mid-span is the reflection of the bottom of the aircraft. The bright area near the top is the forward propeller and forward engine exhaust reflections. Achieving an N-factor greater than 14 with the polished leading edge demonstrates the low-turbulence environment of flight. Results such as these have never been obtained in wind tunnels where N-factors of 8-9 have been achieved with N = 6 being more common. With 80% laminar flow, there is not much that can be done with DRE for laminar flow control. However, the polished leading edge with 0.33 m rms can be considered a base state. A more realistic, operational surface would be painted. D. Laminarization results with a painted surface The model surface was painted to achieve a background roughness level of 1.0 m rms with a 3.8 m avg peakto-peak. In this case transition moved forward to 25% to 30% chord under conditions of Rec = 8.0 x 106, AoA = -4, = 30 and an N-factor = 7. This is shown in figure 3.

Figure 3. IR thermography at 173 KTAS, AoA = -4, Rec = 8.0 x 106, no DRE, White painted LE. x/c)tr 30%. In this case transition moved forward to 30% chord and this is our new base state. When a double layer of DREs (12 m high) was used, the transition location moved back to 60% chord. Rec = 8.0 x 106, AoA = -4, = 30 and an N-factor = 12. This shown in Figure 4. The region of laminar flow was doubled from the base state and, according to linear theory, the disturbance amplitude was reduced by e-7 or < 10-3. This rather remarkable result demonstrates the periodic DRE technology in flight at a chord Reynolds number of 8 million.

6 American Institute of Aeronautics and Astronautics


092407

Figure 4. 180 KTAS, AoA = 0, Rec = 8.13 x 106, White painted LE, DRE x 2 placed at 1% x/c at inboard pressure row, and 1.3% x/c at outboard pressure row, d = 1 mm, = 2.25 mm, transition moved to 60% x/c

Acknowledgments
The work has been supported by the Air Force Research Laboratories, WPAFB, the Air Force Office of Scientific Research, and the Northrop-Grumman Corporation. The authors acknowledge the supporting efforts of the pilots: M.L. Roy Martin, Dr. Donald Ward, and Celine Kluzek and the ground crew: Jerrod Hofferth, Lauren Hunt, Cecil Rhodes, and Shane Schouten.

References
Balakumar P, Reed HL. 1991 Stability of three-dimensional supersonic boundary layers. Phys. Fluids A, 3(4):617. Barker SJ, Gile D. 1981 Experiments on heat stabilized laminar boundary layers in water. J. Fluid Mech. 14:139-58 Bippes H, and Nitschke-Kowsky P. 1990 Experimental study of instability modes in a three-dimensional boundary layer. AIAA J. 28(10):1758-63 Bippes H, Mller B, Wagner M. 1991 Measurements and stability calculations of the disturbance growth in an unstable three-dimensional boundary layer. Phys. Fluids 3(10):2371-7 Bippes H. 1997 Environmental conditions and transition prediction in 3-D boundary layers. AIAA Paper No.97-1906 Boeing Commercial Airplane Group. 1993 High Reynolds Number Hybrid Laminar Flow Control (HLFC) Flight Experiment Final Report. Volume I: D6-55648-1 through Volume 6: D6-55628-6 Bushnell DM. 1992 Overview of aircraft drag reduction technology. AGARD Report 786 Special course on skin friction drag reduction Bushnell DM, and Tuttle MH. 1979 Survey and bibliography on attainment of laminar flow control in air using pressure gradient and suction. Volume I. NASA RP-1035. Dagenhart JR, Saric,WS, Mousseux MC, Stack JP. 1989 Crossflow-vortex instability and transition on a 45-degree swept wing. AIAA Paper No. 89-1892 Dagenhart JR, Saric WS, Mousseux MC. 1990 Experiments on swept-wing boundary layers. Laminar-Turbulent Transition III. Eds. D. Arnal, R. Michel, Springer Deyhle H, Bippes H. 1996 Disturbance growth in an unstable three-dimensional boundary layer and its dependence on initial conditions. J. Fluid Mech. 316:73113 7 American Institute of Aeronautics and Astronautics
092407

Deyhle H, Hhler G, Bippes H. 1993 Experimental investigation of instability wave propagation in a 3-D boundarylayer flow. AIAA J. 31(4):637-45 Dougherty NS Jr, Fisher DF. 1980 Boundary-layer transition on a 10 cone. AIAA Paper No. 80-0154 Haynes TS, Reed HL. 2000 Simulation of swept-wing vortices using nonlinear parabolized stability equations. J. Fluid Mech. 405:325-49 Joslin RD 1998 Laminar Flow Control. Ann. Rev. Fluid Mech. 23 Kohama Y, Saric WS, Hoos JA. 1991 A high-frequency secondary instability of crossflow vortices that leads to transition. Proc. R.A.S. Boundary Layer Transition and Control Liepmann HW, Nosenchuck DM. 1982 Active control of laminar-turbulent transition. J. Fluid Mech. 118:201 Lowell RL, Reshotko E. 1974 Numerical Study of the Stability of a Heated Water Boundary Layer. Report FTAS/TR-73-93, Case Western Reserve University Mack LM. 1984 Boundary-layer linear stability theory. AGARD Report No. 709 Special course on stability and transition of laminar flows Maddalon DV, Collier FS, Montoya LC, Putnam RJ. 1990 Transition flight experiments on a swept wing with suction. Laminar-Turbulent Transition III eds: D. Arnal, R. Michel, Springer Verlag, New York Malik MR, Li F, Chang CL. 1994 Crossflow disturbances in three-dimensional boundary layers: Nonlinear development, wave interaction and secondary instability. J. Fluid Mech. 268:1-36 Matishek J. 2007 Introduction to the Aerion Supersonic Business Jet. Canadian Aero. Space Inst. Symposium Pfenninger W. 1977 Laminar flow control - Laminarization. AGARD Report No. 654, Special Course on Drag Reduction Pupator PT, Saric WS. 1989 Control of random disturbances in a boundary layer. AIAA Paper No. 89-1007 Radeztsky RH Jr, Reibert MS, Saric WS. 1994 Development of stationary crossflow vortices on a swept wing. AIAA Paper No. 94-2373 Reed HL, Nayfeh AH. 1986. Numerical-perturbation technique for stability of a flat-plate boundary layer with suction. AIAA J. 24:208. Reed HL, Balakumar P. 1990 Compressible boundary-layer stability theory. Phys. Fluids A 2(8):1341 Reibert MS, Saric WS, Carrillo RB Jr, Chapman KL. 1996 Experiments in nonlinear saturation of stationary crossflow vortices in a swept-wing boundary layer. AIAA Paper No. 96-0184 Reshotko E. 1978 Heated boundary layers. Proc. 12th Symposium on Naval Hydrodynamics, Nat'l Acad. Sci. Reshotko E. 1979 Drag Reduction by Cooling in Hydrogen-Fueled Aircraft. J. Aircraft 16(9):584-90 Reshotko E. 1984 Laminar flow control - Viscous simulation. AGARD Rpt. 709, Special Course on Stability and Transition of Laminar Flows Reshotko E. 1985 Control of boundary-layer transition. AIAA Paper No. 85-0562 Reynolds GA, Saric WS. 1986 Experiments on the stability of a flat-plate boundary layer with suction. AIAA J. 24:202 Saric WS, Nayfeh AH. 1977 Nonparallel stability of boundary layers with pressure gradients and suction. AGARD C-P No. 224 Saric WS. 1985a Boundary-Layer Transition: T-S Waves and Crossflow Mechanisms. AGARD Report No. 723. Special course on aircraft drag prediction and reduction Saric WS. 1985b Laminar Flow Control with Suction: Theory and Experiment. AGARD Report No. 723. Special course on aircraft drag prediction and reduction Saric WS. 1990 Low-speed experiments: Requirements for stability measurements. Instability and Transition I. Ed. Y. Hussaini, Springer-Verlag Saric WS, Carrillo RB Jr, Reibert MS. 1998 Nonlinear stability and transition in 3-D boundary layers. Meccanica 33:469-87 Saric WS, Reed HL, White EB. 2003 Stability and Transition of Three-Dimensional Boundary Layers. Ann. Rev. Fluid Mech. 35:413-40 Strazisar AJ, Reshotko E, Prahl JM. 1977 Experimental study of the stability of heated flat plate boundary layers in water. J. Fluid Mech. 83(2):225-47 8 American Institute of Aeronautics and Astronautics
092407

Sturdza P. 2007 Extensive supersonic natural laminar flow on the Aerion business jet. AIAA Paper No. 2007-0685 Thomas ASW. 1983 Control of boundary-layer transition using a wave superposition principle. J. Fluid Mech. 137:233 Wagner RD, Maddalon DV, Fischer DF. 1988 Laminar flow control leading edge systems in simulated airline service. ICAS-88.3.7.4 Wassermann P, Kloker M. 2002 Mechanisms and control of crossflow-vortex induced transition in a threedimensional boundary layer. J. Fluid Mech. 456:49-84 White EB, Saric WS. 2005 Secondary instability of crossflow vortices. J. Fluid Mech. 525:275-308

9 American Institute of Aeronautics and Astronautics


092407

Vous aimerez peut-être aussi