You are on page 1of 12

AIAA 2001-0124 Application of MEMS Devices to Delta Wing Aircraft: From Concept Development to Transonic Flight Test

A.Huang, C. Folk, C. Silva, B. Christensen, Y. Chen, an d C.M. Ho UCLA Mechanical and Aerospace Department Los Angeles, CA F. Jiang, C. Grosjean, and Y.C. Tai California Institute of Technology Pasadena, CA G.B. Lee National Cheng Kung University Tainan, Taiwan M. Chen Feiya Technology Corporation Taipei, Taiwan S. Newbern AeroVironment, Inc. Simi Valley, CA

AIAA-2001-0124 APPLICATIONS OF MEMS DEVICES TO DELTA WING AIRCRAFT: FROM CONCEPT DEVELOPMENT TO TRANSONIC FLIGHT TEST
Adam Huang*, Chris Folk*, Chris Silva*, Brian Christensen*, Yih-Far Chen*, Gwo-Bin Lee, Minjdar Chen, Scott Newbern, Fukang Jiang , Charles Grosjean , Chih-Ming Ho*, and Yu-Chong Tai
Abstract About six years ago, we conceived the idea of using Micro-Electro-Mechanical Systems (MEMS) based transducer for aircraft maneuvering. By using micro-actuators as effectors, micro-sensors to detect the optimal actuation location, and microelectronics to provide close loop feedback decisions, a low power control system has been developed for controlling a UAV-sized aircraft. This paper covers the evolution of developing a MEMS controlled delta wing UAV (christened Gryphon), from concept to application, and the transonic testing of the sensors and actuators. Furthermore, recent concepts have been developed to further simplify and add robustness to the control scheme (Please refer to AIAA 20010121). Introduction It can be safely assumed that vortex control follows almost immediately after the discovery of leading-edge vortex flow in 1946 at NASA Langley Research Center1 . However, early researches (195060s) by leading aerodynamic institutions focused on using the delta wing planform to counter compressibility effects. Although, the delta wing planform is almost synonymous with leading edge vortex flow, the original intention was to delay the drag due to compressibility by using sweptback and sharp leading edges. This invariably led to the desire to control the vortex into symmetric pairs and enhancing L/D ratio. Although, the leading edge vortex flow offers higher maximum lift coefficient, CL,max , a higher drag is noticeable at subsonic conditions where the induced drag portion of the total drag is dominant. _________ *University of California, Los Angeles, California National Cheng Kung University, Tainan, Taiwan Feiya Technology Corporation, Taipei, Taiwan AeroVironment, Inc., Simi Valley, CA 93064, USA California Institute of Technology, California Copyright 2000 The American Institute of Aeronautics and Astronautics Inc. All rights reserved. To be more exact, the L/D ratio is actually lower than 2-D airfoils in subsonic flights since the leading edge suction provided by the potential flow is decreased by the creation of the vortex flow with sharp leading edges. Hence, the desire of increasing L/D at large Mach number range (subsonic to supersonic) by using various leading edge shaping and span-wise geometry resulted with the advanced and complex wing design of the Anglo-French Concorde. The requirement of high performance lightweight fighters in the late 1960s lead to the renewed interests in leading edge vortex flows. By using leading edge root extensions (LERXs), a significant amount of wing lift can be generated/concentrated by the highly swept root sections while the majority of the wing with lower sweep angles remains low drag (before stalls) and low structural stress. Again, the vortex control method used is static through tailoring LERX and wing leading edge geometries. This time the desire was to enhance the maximum lift and vortex pair stability at high Angles-ofAttack (AOA) situations as required by the size/weight limit for the lightweight fighter designs. Figure 1 clearly shows the formation of two vortices over the LERXs created by an US Navy Blue Angel in high AOA pullout maneuvers.

(US Navy) Figure 1: Modern fighters rely on LERXs to generate high lift vortices (F/A-18). Following the development of YF-16 and YF-17 (later modified as USN F-18), NASA and US Air Force aerodynamic researches begin interested in the notion of leading edge vortex control in the form of vortex flaps. The primary emphasis was to seek a high L/D solution in subsonic through supersonic region in order to support the requirement of super-cruise flight for the next generation of tactical fighter aircraft 1 . The idea behind

1 American Institute of Aeronautics and Astronautics

the vortex flap is to trap the vortex along the leading edge section of the wing to regain the previously lost vortex suction when fixed sharp leading edges are used. Instead of large vortex structures convecting downstream along the chord axis direction of the leeward side, the vortex is made to stay near the leading edge along the entire span of the high sweep areas. Thus, the flap itself acts as the surface boundary for the vortex to provide a net suction thrust on the aircraft 2 (see Figure 2).

and up (even todays advanced millimeter wave radars are about one wavelength in length scale with the developed MEMS actuators used in this research). The project took on the challenge to develop the required MEMS sensor and actuators that were not available at the time while dealing with the theoretical feasibility of using such small length scale actuators to control a large-scale machine. This paper describes the research that leads to todays distributed MEMS-controlled UAV called Gryphon. Figure 3 illustrates the conceptual drawing of an aircraft controlled by micro-sensors, -actuators, and electronics. This type of closed loop feedback system is called the M 3 microsystem37,38.

Figure 2: The vortex flap. The leading edge vortex is limited to the flap section2 . It is clear from the theory of vortex flap that by cleverly using differential affects to cause loading asymmetry, control forces and moments for all three axis about the aircraft can be created. Thus, Rao3 postulated on the usage of a modified flap called vortex cavity flap to create roll, pitch, and yaw controls while offering drag reduction. In addition to physical flap methods, the following techniques to control leading edge vortex flow has been tried by numerous researchers: a) blowing4-17 , b) suction18-19 , c) trailing-edge jet control20-21 , and d) heating22 . A new method of leading edge vortex control was envisioned in 1994 by using Micro-ElectroMechanical System (MEMS). The concept was to provide distributed control of the leading edge vortex flow through tens if not hundreds of millimeter/micron sized actuators. This was believed not only to provide the type of maneuvering postulated by Rao3 , but by being better, faster, and spending less power. These are the intrinsic properties of typical MEMS devices and benefits of distributed control logic schemes. The size of the micro effectors also serve as a solution to observablility problems by noting that most radars use wavelengths on the order of a few centimeters

Figure 3: The conceptual drawing of MEMS controlled UAV with M 3 microsystem. Vortex Shifting Concept This project involves demonstrating the concept of using small actuators (~micro-millimeter scale) to provide large control forces for a large-scale system (~meters scale). This was theorized by using fluid separation phenomenon, vortex symmetry, and vortex dynamics on a delta wing aircraft. It has been known that the ensuing leading edge vortex flow is extremely sensitive at the separation location of the free shear layer on a swept aerodynamic planform. Sharp leading edges have a natural boundary condition of separation at the sharp edges. For round leading edges, the genesis location of the free shear layer is a complex curved line that runs along the leading edge from the apex to the wing tip. Changing the surface boundary condition in quasi-steady state (i.e., expanding surface curvature) near the separation line can cause a global change to the developing vortex. Lee, et al.23 has compiled an excellent comprehensive review on the topic of lift forces of delta wings. Furthermore, it has been shown that instability effects can also manipulate separated vortex flows. Ho, et al.24 has shown that very small forcing amplitudes can control the vortex structures downstream of the mixing layer if the correct sub-harmonic perturbation of the most

2 American Institute of Aeronautics and Astronautics

energetic frequency is used to manipulate the spreading rate in the vortex evolution. In general, delta wing leading edges create symmetric pairs of primary and secondary vortices. The controlled separation of the thin boundary layer at the synthesis location of the vortex pairs on a delta wing aircraft will allow the manipulation of pressure field about the aircraft. By using micro actuators along the round leading edges to create vortex shifts, a resulting loading change about the aircraft would occur. Although at first glance this appears to be the mere adoption of vortex flap with MEMS, however, the difference lies in the fact that the vortex being controlled is not trapped at the leading edge. The round leading edge allows the manipulation of the vortex pairs on the entire leeward side of the aerodynamic surfaces. The length scale of the micro actuators will need to be compatible with the boundary layer thickness. It was speculated from the beginning that the boundary thickness would be the key to the flap/vortex interaction. The boundary thickness was estimated to be between 100s of microns to 1-2 millimeter, depending on the length scale of the aircraft and the Reynolds number. Figure 4 shows that pairs of vortex system increase size and strength with increasing AOA.

Thus, it was hypothesized that the vortex shifting mechanisms can be used to replace or supplement conventional surfaces at high angles of attack where the latter became ineffective due to trailing edge separation of the potential flow. Actuators The first problem is to find a suitable MEMS actuator design for this project. It can be said that even today large force and displacement actuators with sufficient frequency response of ~10-100mN, ~1mm, and >10hz, respectively, are hard to come by. Initially, a mechanical actuator using 1-mil stainless steel sheets was used to simulate the MEMS actuators 25,26. It should be noted that aerodynamically, there are no differences between the stainless steel actuator and flap type MEMS actuators. A servomotor was used to actuate the stainless steel strips to actuator lengths of 1 -5mm. Figure 5 shows the wind tunnel model used for the stainless steel actuators. Like all the subsequent experiments, a diameter rod was used as the leading edge of the flat-plate type wind tunnel model.

Figure 5: actuators.

The stainless steel flap type leading edge

Werle, 1958 Figure 4: Typical evolution of primary vortex pairs with increasing AOA on delta wings.

Next, the permalloy magnetic flap actuators were used. The MEMS micro flap is a surface-micromachined plate actuated by an external permanent-magnet mechanism. The 1 mm x 2mm actuator was fabricated from electroplating solution that yields the electroplated Ni80 Fe20 alloy27,28. The Ni80 Fe20 film was used to induce magnetic forces from an external permanent magnet source. Although MEMS electromagnetic coils have been produced in the past, the amount of loading estimated for this project was calculated to be too high. This would require a high electric current on the orders of amps for the required actuation heights and forces 29 . Figure 6 shows the MEMS permalloy and electro-magnetic actuator fabricated for the delta wing wind tunnel model.

3 American Institute of Aeronautics and Astronautics

current MEMS balloon actuator is fabricated on metal substrate (brass or copper). The 50-mil copper substrate used on the final version allows a conformal surface to be covered by the flexible balloon actuators. Figure 8 shows the flexible m etal substrate wrapped around a leading edge rod used for the wind tunnel model.

(a)

(b)

Figure 6: (a) MEMS permalloy flap type micro actuator. (b) MEMS electromagnetic coil actuators. It is rather obvious that the MEMS flap type actuators will not be the solution to this project. MEMS actuators in general has the capability of producing <100uN30 , which is at least 1 order of magnitude off from the requirement. Also, MEMS flap actuators are not robust enough for high-speed flights. It has been demonstrated in the wind tunnel that at speeds higher than 30m/s, a large percentage of the MEMS leading edge flap actuators were stripped off by the free-stream. A new type of actuator called bubble actuator was selected as the definitive effectors for this project. The bubble actuator utilizes an external pneumatic source to inflate a spin-on silicone rubber membrane. This type of actuator is theoretically able to produce extremely high forces (>10mN) since the actuation principle relies on the pressure drop between the balloon membranes (>5 psi is easily achievable). The MEMS bubble actuators had gone through 2-3 generations. The first generation used anisotropically etched silicon substrates to provide the manifold for gas injection. This type of actuator was designed for ease of fabrication on existing silicon wafer processes 31 . Figure 7 shows the silicon substrate bubble actuators being inflated at various pressures.

Figure 8: actuators.

Flexible MEMS metal substrate bubble Sensors

Figure 7: Silicon substrate MEMS bubble actuators. However, the rigid substrate is not a practical solution for adoption on aircraft surfaces. The

The primary sensor used for this project is the MEMS shear stress sensor. It was very clear since the onset of this project that a high sensitivity and robust shear stress sensor is needed in order to detect the sensitive location of the curved separation line. Originally, a flap type actuator on the sharp leading edge was also considered since this would negate the requirement of shear stress sensors. However, due to MEMS actuator power constraints and geometrical limitations this part of the project was postponed until suitable actuators are available. Recently, preliminary tests were done using mechanical strips on the leading edge of sharp actuators to simulate a micro flap. Please refer to AIAA 2001-0121 for a dedicated report on this topic. Earlier actuator tests were done in a systematic fashion of eliciting the control forces and moments. This fashion of blind open loop actuation was extremely time consuming and ineffective. For example, the largest rolling moment coefficient created by the MEMS flap actuators was found to be only on the order of ~0.001. The availability of shear stress sensor quickly allowed the authors to target the region near the separation line. Luckily, the rolling moment coefficient was later found to be more than 1 order of magnitude higher with the correct actuation location. However, the sensor technology didnt come easily. At the beginning of this research, MEMS pressure and temperature sensors were already rather mature. This is

4 American Institute of Aeronautics and Astronautics

due to the wide spread requirement for this kind of sensors in different disciplines, and they also have high commercial values. However, shear stress sensors are generally only useful to applied fluid- and aero-dynamicists. Thus, there were only a few designs available. Moreover, measuring surface shear forces are not as straightforward as pressure or temperature. Traditional transducers such as Stanton32,33 and Preston34,35 tubes are either too large or has slow frequency response. This project requires high response (>100hz) and high spatial resolution shear stress sensors for close loop feedback purposes. In terms of MEMS shear stress sensors, there are various methods of measuring shear stress36 (i.e., piezoresistive, capacitive, optical, mechanical tether, and thermal). The thermal type shear stress sensor was selected for this work since experiences were gained through the fabrication of MEMS hot wire anemometers 39 . The first MEMS thermal shear stress sensor was fabricated on silicon substrates. Later, an array of sensors acting as a shear stress imager was used for turbulent shear streak experiments in 1D turbulent wind tunnels 40-43 . However, the rigid substrate shear stress sensor is not a practical solution for the round leading edge of a delta wing aircraft. Early experiments using rigid silicon substrates required a relentless dedication to measure thousands of leading locations by rotating the sensors at each angle of the leading edge rod and manually displacing along the leading edge. Figure 9 shows the MEMS shear stress imager with the magnification of each sensor element. The imagine zone is about 1cm x 1cm and contains over 100 shear stress sensors or temperature sensors (including test sensors). Each element area is about 250um x 250um and the sensing hot bridge is a 150 um x 3 um x 0.5 um polysilicon wire. Theoretically, a much finer resolution of < 50 um x50 um element area is possible if the interconnecting aluminum wires are more efficiently routed or by using an on-chip CMOS circuitry to distribute power and process the data into digital format.

Figure 9: The silicon substrate MEMS shear stress imaging sensor chip and magnification of each sensor element. Like the actuators, a flexible skin solution must be achieved for realistic usage on the curved surfaces of round leading edge delta wings. The flexible skin MEMS shear stress sensor was indeed developed specifically for this project. By using thin polyimide films, a flexible membrane holding each shear stress sensor in a faceted manner (silicon islands) was fabricated. There are roughly four iterations o design for this flexible skin f MEMS shear stress sensor and is currently extremely mature. Figure 10 shows the latest generation of flexible shear stress sensor and the associated flexible Kapton PCB. After ironing on the flexible MEMS shear stress skin to the PCB, the PCB is then wrapped on a small section of the leading edge radius. This sensor array contains 36 shear stress sensor elements coated with a thin layer of silicon nitride for maximum robustness in harsh environments 44 .

Figure 10: Third generation flexible MEMS shear stress sensor array with flexible PCB. The resulting flexible unit with the PCB is then wrapped around the desired leading edge section.

5 American Institute of Aeronautics and Astronautics

Wind Tunnel Experiments The wind tunnel experiments for this project can be roughly broken down to five sections based primarily on the availability of the MEMS sensors and actuators. This was found to be the toughest part of this research since all the MEMS sensors and actuators need to be designed, fabricated, and tested in-house. Although currently better understood, MEMS design and fabrication is likened to be blackart since any lack of experience in this field will dramatically limit the yield of the devices fabricated (often resulting in a yield of zero). Rolling Motion Demonstration The first experiment performed by the authors to show the feasibility of this concept was in 1994. An aluminum delta wing wind tunnel model was mounted on a low friction bearing. By using servomotors to actuate a small piece of stainless steel strip near the leading edge on the windward side (lower surface), rolling motion of the wind tunnel model was observed. Both DC and AC actuations (Figure 11 &12) were used and the data consistently shows a rotation of ~12-15 degrees of roll45 .

from figure 11 and 12 that the difference between 2 and 3mm actuation is minimal. This indicates that the boundary thickness is roughly 2mm and supports the original hypothesis that the local boundary thickness is an important length scale to be considered for the actuation length.

Figure 12: Unsteady actuation of the mechanical leading edges at 0.5 Hz. Flap Actuators On Round Leading Edges After the first free rolling motion was observed, the next plan was to quantify the amount of rolling moment coefficients and to test the possibility of producing other forces and moments useful for aerodynamics. The wind tunnel model is mounted on a six-component force balance that is connected to a PC-controlled ADC card. The bulk of the data during this phase uses the stainless steel strips mentioned in the actuator section. Later, with the availability of MEMS flap type magnetic actuators, the experiment was repeated to show that the results are similar. This is not surprising since the boundary condition provided by the actuators is invariant of materials with similar geometrical rigidity. Figure 13 shows the rolling moment coefficients created by the 2mm stainless steel leading edge actuators. Percentage change of rolling moment is normalized by the vortex moment, which was assumed to be the lift force47 produced by a single vortex multiplied by the distance from the centerline to the centroid of the semi-span wing. The beta angle, , is measured from the bottom surface to the top surface (0-180 degrees)25,46.

Figure 11: (a) Excitation signal for the mechanical leading edge actuators. (b)-(d) Varying the actuator flap length. The wind tunnel model used was a 5 mid-chord 56.5 degrees delta wing. It was during this phase of the research that the swept angle used for the most subsequent wind tunnel models was based on the inner wing section of the X-31, 56.5 degrees. Notice

6 American Institute of Aeronautics and Astronautics

position. The beta angle means the same as in previous and all the following cases (0-180 degrees). The sensor resolution on the 0.5 leading edge rod is limited to 6 degrees (30 sensors). The boundary line of the dark and light regions indicates dramatic shear stress change, which was defined as the critical separation line. This is merely the reference line for the actuators to be positioned for effective vortex shifts 25 . Surface Pressure Measurement and Smoke Particle Flow Visualization. Although, MEMS actuators created the recorded forces/moments, a clearer understanding of the actual physical phenomenon was needed as a sanity check. The entire starboard section of the wind tunnel was tapped with over 120 pressure taps. Surface pressure measurements indicated that the MEMS actuation causes the shift in the peak suction areas. This is indicative of a vortex shift as described by the theory. Surface integration was performed to yield similar results collected by the force balance. However, it was felt that a visual check was need to explain the actual phenomena since the pressure measurement only indicated the normal surface force. A pulsed Nd-Yag laser was used to spread a thin laser sheet over the leeward side of the wind tunnel model. The stainless steel actuators were used in this case since the oil from the smoke generator can efficiently destroy the MEMS flap actuators. Indeed, the vortex was found to be shifting (Figure 15) with the actuators, and frequency response up to 10 hertz was followed by the vortex shifts (the 10 hertz limit was due to the servo drive). An interesting note should be made that even the burst portion of the vortex was found to be shifting. Furthermore, the actuators can twist the vortex convection path. What this means is that a reversal in shifting direction was found in some cases. For exa mple, the primary vortex near the apex region shifts inboard while the trailing edge region shifts outboard.

Figure 13: The normalized rolling moment produced by the flap type leading edge actuators. Flap Actuators with Shear Stress Information With the availability of the MEMS shear stress sensors, the authors were able to find an order of magnitude higher in rolling moment. Furthermore, the MEMS sensors provided important clues for demonstrating control forces/moments in other axis. Figure 14 shows the DC shear stress level produced by the shear stress sensor.

Figure 14: Leading edge shear stress map from the wind tunnel model. The data was taken by shear stress sensor arrays staggered along the starboard leading edge. The position from apex was shifted 3 cm along the leading edge (physical constraint limited the sensor at the apex area) and the stagger increments at 1cm per

Figure 15: Smoke particle laser-sheet flow visualization25 .

7 American Institute of Aeronautics and Astronautics

MEMS Bubble Actuator and Flexible MEMS Shear Stress Sensors The availability of the MEMS bubble actuators made this project more feasible. The MEMS bubble actuators provides the robustness and high forces required by real airframes. Unfortunately, the control forces provided by the bubble actuators yielded different force/moment causality. The effects from the bubble actuator were found to be extremely sensitive to the Reynolds number. Like the analogy between the sharp and round leading edge delta wings, this was partly anticipated due to the smooth bubble actuators as compared to the sharp-edged flap type actuators. A comparison between the bubble actuators and conventional aileron surfaces were performed. The conventional aileron used (outer portion, of the semi-span) has a chord of 19.4% of the mean aerodynamic chord while the MEMS bubble actuator is only 0.8% of the mean aerodynamic chord. Figure 16 shows the rolling moment comparison between these two types of control surfaces. The aileron deflections used were +20/-25 degrees. It should be noted that the MEMS bubble actuators used were linear, this means that Figure 16 may not show the maximum possible by the bubble actuators. As mentioned before, the rolling moment produced is extremely sensitive to the actuator location, thus this part of the experiment is still ongoing in search of the highest moment achievable.

some software fixes as experience in the M3 system is gained in the wind tunnel. The final tests demonstrated the ability of the DSP being able to pick the correct actuating location based on the shear stress sensor data. The entire wind tunnel mo del with 60 sensors and 12 linear array flexible bubble actuators were tested on the starboard side. Figure 17 shows the M 3 system inside the wind tunnel.

Figure 17: M3 system on wind tunnel model with free rolling axis. Rotations were detected with the M 3 system. Flight Tests As the results from the wind tunnel experiments begin to build confidence into the theory of vortex manipulation by MEMS actuators, flight tests of sensors and actuators were sought after for the ultimate challenge. Initially, a 1/7th scale radio controlled ducted-fan Mirage III model was used to demonstrate the actuator effects in the real world. The Mirage flights utilized the same stainless steel mechanical actuators mounted in the wind tunnel. Many maneuvers were achieved using the Mirage III model, such as 270-degree turns and barrel rolls. This was the first demonstrated flight of vortex shifting with our leading edge micro actuator concept (Figure 18).

Figure 16: The comparison between conventional aileron surfaces and MEMS bubble actuators. M3 System in Wind Tunnel Prior to be fitted into a flying research aircraft, the entire system was fitted into the wind tunnel for component and system checks. A DSP was used to control the data acquisition of the shear stress sensors, processing of the sensor data, and actuation logic. The entire system performed as designed after

Figure 18: The Mirage III radio-controlled demonstrator for vortex shift concept.

flight

8 American Institute of Aeronautics and Astronautics

After the successful demonstration by the Mirage III, a dedicated research aircraft with full onboard instrumentation and data acquisition system was designed to quantify the M3 system in real world conditions. This aircraft was christened the Gryphon UAV and the whole project was affectionately nicknamed the Gryphon UAV Project.

Figure 21: MEMS shear stress sensor and bubble actuators on NASA Dryden F-15B. Conclusion An entire M system has been developed to provide a MEMS-based control mechanism for a delta wing aircraft. The aerodynamics of vortex shifting with actuators on the order of boundary thickness has been demonstrated. New flexible sensor and actuator technology has been produced using micro-machining techniques. Currently, we are at the last phase of this project and the definitive flight research MEMS UAV M3 aircraft has been completed. This new aircraft is named Super Gryphon and encompasses the lessons learned from Gryphon. In addition, an onboard GPS was installed to permit real autonomous flights if such demonstration is deemed necessary. At the time this report was written, the entire M3 system was undergoing shake down before the first flight of the close loop controlled MEMS flight. Hopefully, this project can contribute another success story to the history of aerodynamics. Figure 22 shows the Super Gryphon with the M 3 System being installed. Acknowledgements The authors would like to thank the Defense Advanced Research Projects Agency (DARPA) and the NASA DRYDEN Research Flight Center for the support of the projects. We also like to thank numerous individuals in this large project covering over 6 years.
3

Figure 19: The fully instrumented MEMS research aircraft, Gryphon UAV. AeroVironment, Inc. was tasked with designing and maintaining the Gryphon UAV. The Gryphon UAV contained one acceleration and one rate sensor per primary axis (body axis). It also contained a nose mounted Pitot Tube with AOA and side-slip vanes. A color video camera was mounted on the fin top and real time downlink was used for the video and pertinent flight information such as AOA and flight velocity. The rest of the data was collected with onboard PCMCIA memory cards with an onboard flight processor (TattleTale-8TM). In addition to the Gryphon flight tests, a contract was provided by NASA Dryden to test flight the developed MEMS devices on the Dryden F-15B. Figure 20 shows the mounting of the MEMS shear stress sensor and bubble actuators on the root section of the starboard side of the main wing. In addition, the sensor data was collected to flight speeds up to mach 1.5 and all devices survived the test flight. This truly demonstrated the robustness and practicality of using MEMS aerodynamic devices for flight vehicles.

9 American Institute of Aeronautics and Astronautics

Figure 22: The Super Gryphon with 100 MEMS shear stress sensors (on 5 arrays) and 24 linear MEMS bubble actuators being installed. Bibliography [1] Polhamus, E.C. Vortex Lift Research: Early Contributions and Some Current Challenges, Vortex Flow Aerodynamics, NASA CP2416. pp. 1-30. [2] Campbell, J.F., Osborn, R.F., Leading-Edge Vortex Research: Some Nonplanar Concepts and Current Challenges, Vortex Flow Aerodynamics, NASA CP2416. pp. 31-63. [3] Rao, D.M. Towards An Advanced Vortex Flap System-the Cavity Flap, Vortex Flow Aerodynamics, NASA CP2416. pp. 219-230. [4] Bradley, R.G. and Wray, W.O., A Conceptual Study of Leading-Edge-Vortex Enhancement by Blowing, Journal of Aircraft, Vol. Aa, No. 1, 1974, pp.33-38. [5] Wood, N.J., and Roberts, L., The Control of Vortical Lift on Delta Wings by Tangential Leading Edge Blowing, AIAA paper 87-0158, Jan 1987. [6] Campbell, J.F., Augmentation of Vortex Lift by Spanwise Blowing, Journal of Aircraft, Vol. 13, No. 9, 1976, pp.727-732.

[7] Greenwell, D.I., and Wood, N.J., Roll Moment Characteristics of Asymmetric Tangential Leading-Edge Blowing on a Delta Wing, Journal of Aircraft, Vol. 31, No. 1, 1994, pp.161-168. [8] Johari, H., Olinger, D.J., and Fitzpatrick, K.C., Delta Wing Vortex Control via Recessed Angled Spanwise Blowing, Journal of Aircraft, Vol. 32, No. 4, 1995, pp. 804-810. [9] Gu, W., Robinson, O., and Rockwell, D., Control of Vortices on a Delta Wing by Leading-Edge Injection, AIAA Journal, Vol. 31, No. 7, 1993, pp.1177-1186. [10] Miyaji, K., Fujii, K., and Karashima, K., Enhancement of the Leading-Edge Separation Vortices by Trailing-Edge Lateral Blowing, AIAA Journal, Vol. 34, No. 9, 1996, pp.1943-1945. [11] Meyer, J., and Seginer, A., Pulsating Spanwise Blowing on a Fighter Aircraft, AIAA Paper 92-4359, 1992, pp.188-197. [12] 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. 1024-1049. [13] Celik, Z.Z., and Roberts, L., Aircraft Control at High-Alpha by Tangential Blowing, AIAA Paper 920021, Jan. 1992 [14] Greenwell, D.I., and Wood, N.J., Static Roll Moment Characteristics of Asymmetric Tangential Leading Edge Blowing on a Delta Wing at High Angles of Attack, AIAA Paper 93-0052, Jan. 1993. [15] Fitzpatrick, K., Johari, H., and Olinger, D., A Visual Study of Recessed Angled Spanwise Blowing Method on a Delta Wing, AIAA Paper 93-3246, July 1993. [16] Malcom, G.N., and Skow, A.M., Flow Visualization Study of Vortex Manipulation of Fighter Configurations at High Angle of Attack, CP-413 AGARD, 1986. [17] Alexander, M., and Meyn, L.A., Wind Tunnel Results of Pneumatic Forebody Vortex Control Using Rectangular Slots on a Chined Forebody, AIAA Paper 94-1854, June 1994. [18] Hummel, D., Zur Umstromung Sharfkantiger Shlanker Deltaflugel bei Grossen Anstellwinkeln, Zeitschrift fuer Flugwissenschaften, Vol. 15, No. 10, 1967, pp. 376-385. [19] Ross, F.W., and Kegelman, J.T., Recent Explorations of Leading Edge Vortex Flowfields, NASA CP-3149, Vol. 1, Pt. 1, Oct 19990, pp. 157-172. [20] Shih, C., and Ding, Z., Trailing-Edge Jet Control of Leading-Edge Vortices of a Delta Wing, AIAA Journal, Vol. 32, No. 7, April 1996, pp. 1447-1457. [21] Helin, H.E., and Watry, C.W., Effects of TrailingEdge Jet Entrainment on Delta Wing Vortices, AIAA Journal, Vol. 32, No. 4, April 1994, pp. 802-804. [22] Marchman, J.F., Effect of Heating on Leading Edge Vortices in Subsonic Flow, Journal of Aircraft, Vol. 12, No. 2, 1975, pp. 121-124.

10 American Institute of Aeronautics and Astronautics

[23] Lee, M., Ho, C.M., Lift Forces of Delta Wings, Applied Mechanics Reviews (ASME Book No. AMR077), Vol. 43, No. 9, September, 1990. [24] Ho, C.M., and Huang, L.S., Subharmonics and Vortex Merging in Mixing Layers, Journal of Fluid Mechanics, 119, 1982, pp. 443-473.\ [25] Lee, G.B., Control of a Delta-Wing Aircraft by Using Micromachined Sensors and Actuators, Ph.D. Thesis, Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, U.S.A. (1998). [26] Lee, G.B., and Ho, C. M., " Roll Motion Control of a Delta Wing by LE actuators," Bulletin of the American Physical Society, 49th Annual Meeting of the Division of Fluid Dynamics, Nov. 24-26, 1996, Syracuse, New York. [27] Liu, C., Tsao, T., Tai, Y.C., Leu T.S., Ho, C.M., Tang, W., and Miu, D., Micromachined Permalloy Magnetic Actuator for Delta-Wing Control, Proceedings from IEEE Micro Electro Mechanical Systems, Germany, pp. 332-335 (1995). [28] Liu, C., Tsao, T., Lee, G.B., Leu, T.S., Tai, Y.C., and Ho, C.M., Out-of-Plane Magnetic Actuators with Electroplated Permalloy for Fluid Dynamics Control, Sensors and Actuators A: Physical, 78, Issue: 2-3, 1990. [29] Liu, C., Tsao, T., Tai, Y.C., Ho, C.M., 1994 IEEE Workshop on Micro-Electro-MechanicalSystems, pp. 57-62, 1994. [30] Kovac, G.T.A., Micromachined Transducers Sourcebook , McGraw-Hill, New York (1998). [31] Grosjean, C., Lee, G.B., Hong, W., Tai, Y.C., and Ho, C.M., Micro Ballon Actuators for Aerodynamic Control, IEEE MEMS-98 Workshop, Germany (1998). [32] Hool, J.H., Measurement of Skin Friction Using Surface Tubes, Aircraft Eng., Vol. 28, p. 52, 1956. [33] East, L.F., Measurement of Skin Friction at Low Subsonic Speeds by the Razor-Blade Technique, R&M 3525, Aero. Res. Council. London, 1966. [34] Goldstein, R.J., Fluid Mechanics Measurement. Hemisphere Publishing Co., p. 61, 1983. [35] Preston, J.H., The Determination of Turbulent Skin Friction by Means of Pitot tubes, J.R. Aero. Society., Vol. 58, pp. 109-121, 1953. [36] Huang, A., Ho, C.M., Jiang, F., and Tai, Y.C., MEMS Transducers for Aerodynamics-A Paradigm Shift, AIAA 00-0249, January, 2000. [37] C.M. Ho, P.H. Huang, J. Lew, J.D. Mai, V. Lee, Y.C. Tai, Intelligent System Capable of SensingComputing-Actuating, Keynote Address, 4th International Conference on Intelligent Materials, Society of Non-Traditional Technology. Tokyo, Japan, October 1998.

[38] C.M. Ho, P.H. Huang, J. M. Yang, G.B. Lee, and Y.C. Tai, Active Flow Control by MicroSystems, FLOWCON, International Union of Theoretical and Applied Mechanics (IUTAM) Symposium on Mechanics of Passive and Active Flow Control, Gottingen, Germany, Sept. 1998. pp18-19. [39] Jiang, F., Tai, Y.C., Ho, C.M., Karan, R., and Garstenauer, M., Theoretical and Experimental Studies of Micromachined Hot Wire Anemometers, Technical Digest, International Electron Devices Meeting, San Francisco, CA, pp.139-142, 1994. [40] C. Liu, J. Huang, A. Zhu, F. Jiang, S. Tung, Y.C. Tai, and C.M. Ho, A Micromachined Flow Shear Stress Sensor Based on Thermal Transfer Principles, IEEE/ASME J. of Micro-electro-mechanical Systems (J.MEMS), 1999. [41] F. Jiang, Y.C. Tai, B. Gupta, R. Goodman, S. Tung, J.B. Huang, and C.M. Ho. A Surface-Micromachined Shear Stress Imager, Proceeding, 1996 IEEE Micro Electro Mechanical Systems Workshop (MEMS96), San Diego, pp.110-115, Feb. 11-15 (1996). [42] Huang, J.B., Jiang, F.K., Tai, Y.C., and Ho, C.M., A Micro-Electro-Mechanical-Systems -Based Thermal Shear Stress Sensor with Self-Frequency Compensation, Meas. Sci. Technology, 10, pp. 687-696, 1999. [43] T. Tsao, F. Jiang, R. A. Miller, Y.C. Tai, B. Gupta, R. Goodman, S. Tung, and C. M. Ho, An Integrated MEMS System for Turbulent Boundary Layer Control, Technical Digest, 1997 International Conference on Solid-State Sensors and Actuators (Transducers97), Chicago, IL, Vol. 1, pp. 315-318, June 16-19 (1997). [44] Jiang, F., Xu, Y., Weng, T., Han, Z., Tai, Y.C., Huang, A., Ho, C.M., and Newbern, S., Flexible Sensor Skin for Aerodynamics Applications, MEMS-2000, Miyazaki, Japan, pp. 465-470, 2000. [45] Ho, C.M., Miu, D., and Tai, Y.C., Control of Aerodynamic Devices by Micro Actuation, Proceedings, GOMAC Conference, San Diego, CA, pp. 211-214, Nov, 1994. [46] Lee, G.B., Huang, A., Ho, C.M., Jiang, F., Grosjean, C., Tai, Y.C., Sensing And Control of Aerodynamic Separation by MEMS, The Chinese Journal of Mechanics, Vol. 16, No. 1, March 2000. [47] Polhamus, E.C., Predictions of Vortex-Lift Characteristics by a Leading-Edge-Suction Analogy, Journal of Aircraft, Vol. 8, No. 4, pp.193-199., Vol. 70, No. 5, pp. 420-456, 1971.

11 American Institute of Aeronautics and Astronautics

You might also like