You are on page 1of 55

A Review of Morphing Aircraft

SILVESTRO BARBARINO,1,* ONUR BILGEN,1 RAFIC M. AJAJ,1 MICHAEL I. FRISWELL1 AND DANIEL J. INMAN2
1 2

College of Engineering, Swansea University, Singleton Park, Swansea SA2 8PP, UK

Center for Intelligent Material Systems and Structures, Virginia Tech, Blacksburg, VA 24061, USA ABSTRACT: Aircraft wings are a compromise that allows the aircraft to fly at a range of flight conditions, but the performance at each condition is sub-optimal. The ability of a wing surface to change its geometry during flight has interested researchers and designers over the years as this reduces the design compromises required. Morphing is the short form for metamorphose; however, there is neither an exact definition nor an agreement between the researchers about the type or the extent of the geometrical changes necessary to qualify an aircraft for the title shape morphing. Geometrical parameters that can be affected by morphing solutions can be categorized into: planform alteration (span, sweep, and chord), out-ofplane transformation (twist, dihedral/gull, and span-wise bending), and airfoil adjustment (camber and thickness). Changing the wing shape or geometry is not new. Historically, morphing solutions always led to penalties in terms of cost, complexity, or weight, although in certain circumstances, these were overcome by system-level benefits. The current trend for highly efficient and green aircraft makes such compromises less acceptable, calling for innovative morphing designs able to provide more benefits and fewer drawbacks. Recent developments in smart materials may overcome the limitations and enhance the benefits from existing design solutions. The challenge is to design a structure that is capable of withstanding the prescribed loads, but is also able to change its shape: ideally, there should be no distinction between the structure and the actuation system. The blending of morphing and smart structures in an integrated approach requires multi-disciplinary thinking from the early development, which significantly increases the overall complexity, even at the preliminary design stage. Morphing is a promising enabling technology for the future, next-generation aircraft. However, manufacturers and end users are still too skeptical of the benefits to adopt morphing in the near future. Many developed concepts have a technology readiness level that is still very low. The recent explosive growth of satellite services means that UAVs are the technology of choice for many investigations on wing morphing. This article presents a review of the stateof-the-art on morphing aircraft and focuses on structural, shape-changing morphing concepts for both fixed and rotary wings, with particular reference to active systems. Inflatable solutions have been not considered, and skin issues and challenges are not discussed in detail. Although many interesting concepts have been synthesized, few have progressed to wing tunnel testing, and even fewer have flown. Furthermore, any successful wing morphing system must overcome the weight penalty due to the additional actuation systems. Key Words: bio-inspired, morphing < piezoelectric, shape change, review.

INTRODUCTION

observation of flight in nature has motivated the human desire to fly, and ultimately the development of aircraft. The designs of the first flying machines were relatively crude and even today nature has much to teach us and continuously inspires research. By directly comparing aircraft with nature, designers seek inspiration, in order to achieve the simplicity, elegance, and efficiency that characterize animal species obtained by thousands of years of biological evolution. In particular,
HE

*Author to whom correspondence should be addressed. E-mail: silvestro.barbarino@gmail.com Figures 1 and 3 appear in color online: http://jim.sagepub.com

the attraction for designers is the integration between the structure and function that characterizes the wings of birds (Bowman et al., 2002). Even in complex urban environments, birds are able to rapidly change shape to transition from efficient cruise to aggressive maneuvering and precision descents. Avian morphology permits a wide range of wing configurations, each of which may be used for a particular flight task (Abdulrahim and Lind, 2006). The idea of changing the wing shape or geometry is far from new. The Wright Flyer, the first heavier than air aircraft with an engine, enabled roll control by changing the twist of its wing using cables actuated directly by the pilot. The increasing demand for higher cruise speeds and payloads led to more rigid aircraft

JOURNAL

OF INTELLIGENT

MATERIAL SYSTEMS

AND

STRUCTURES, Vol. 22June 2011

823

1045-389X/11/09 082355 $10.00/0 DOI: 10.1177/1045389X11414084 The Author(s), 2011. Reprints and permissions: http://www.sagepub.co.uk/journalsPermissions.nav

824

S. BARBARINO ET AL.

structures that are unable to adapt to different aerodynamic conditions, characterizing a typical mission profile. The deployment of conventional flaps or slats on a commercial airplane changes the geometry of its wings. The wings are designed as a compromise geometry that allows the aircraft to fly at a range of flight conditions, but the performance at each condition is often sub-optimal. Moreover, these examples of geometry changes are limited, with narrow benefits compared with those that could be obtained from a wing that is inherently deformable and adaptable. The ability of a wing surface to change its geometry during flight has interested researchers and designers over the years: an adaptive wing diminishes the compromises required to insure the operation of the airplane in multiple flight conditions (Stanewsky, 2001). In spite of the apparent complexity of variable-geometry aircraft, nature has evolved thousands of flying insects and birds that routinely perform difficult missions. Observations by experimental biologists reveal that birds such as falcons are able to loiter on-station in a high-aspect ratio configuration using air currents and thermals until they detect their prey. Upon detection, the bird morphs into a strike configuration to swoop down on an unsuspecting prey. Morphing is short for metamorphose and, in the aeronautical field, is adopted to define a set of technologies that increase a vehicles performance by manipulating certain characteristics to better match the vehicle state to the environment and task at hand (Weisshaar, 2006). Using this definition, established technologies such as flaps or retractable landing gear would be considered morphing technologies. However, morphing carries the connotation of radical shape changes or shape changes only possible with near-term or futuristic technologies. There is neither an exact definition nor an agreement between the researchers about the type or the extent of the geometrical changes necessary to qualify an aircraft for the title shape morphing. As early as 1890, the French aviation pioneer Clement Ader proposed a bat wing for an airplane (Eole) that could reduce its size to half or one-third of its full deployment. The first examples of polymorphic aircraft include the Pterodactyl IV, designed by Geoffrey Hill at Westland that flew in 1931, and the Russian fighter IS-1, which flew in 1940 and was able to switch from a maneuverable biplane to a faster monoplane. Weisshaar (2006) gave an historical overview and Figure 1 shows a broad summary of fixed-wing aircraft (of different categories and sizes) implementing flying morphing technologies as a timeline. Historically, morphing solutions always led to penalties in terms of cost, complexity, or weight, although in certain circumstances, these were overcome by the benefits attained at the overall system level. Airplanes such as the F-14 Tomcat and the Panavia Tornado are good examples,

where varying the wing sweep angle allows good performance at both low and high speeds (alleviating the problems of compressibility). Recent developments in SMART materials may overcome these limitations and enhance the benefits from similar design solutions. Chopra (2002) presented a review of the state-of-the-art of smart structures and integrated systems. The design of a morphing aircraft by means of smart materials is a multidisciplinary problem. The challenge is to design a structure that is capable of withstanding not only the prescribed loads, but also to change its shape in order to withstand several load conditions. In order to reduce the complexity and hence increase the reliability, the actuation system, consisting of active materials, should be embedded in the structure. Ideally there should be no distinction between the structure and the actuation system, so that the system used to produce and carry the loads, is also capable of changing shape. In addition to benefits in terms of complexity, reliability, and production cost, such a concept could also prove to be lighter. The timeline in Figure 1 highlights that most large shape modification techniques have been developed for military applications, where a more versatile vehicle compensates for the additional complexity and weight. Furthermore, in recent years, focus has moved to small aircraft (mostly unmanned aerial vehicles, or UAVs). The move toward UAVs results from greater efficiency requirements and a short time-to-deliver because of reduced certification issues and qualification tests. The lower aerodynamic load on UAVs also increases the number of potential morphing technologies. Except for variable-sweep and swing-wing concepts, most previous morphing concepts were applied to lightly loaded, relatively low-speed airplane designs. It is fair to ask why new investment should be made in morphing wings when this concept has been tried in the past and has had so little impact. The answer to this question has two parts, both of which are concerned with the technology that exists today, compared to that which existed several decades ago. First, many new, novel materials, material systems, and actuation devices have been developed during the past decade. These developments allow designers to distribute actuation forces and power optimally and more efficiently. Properly used, these devices reduce weight compared to other, more established designs. Second, missions today differ markedly from those a decade ago. Adversaries may be sophisticated or unsophisticated. Targets are more distributed and are smaller, but the proliferation of sophisticated air defenses mean that these targets may be very dangerous to attack. Morphing provides mission flexibility and versatility to deal with these kinds of targets in a cost-effective manner. Wing shape-changing concepts require actuators attached to internal mechanisms, covered with flexible/

Morphing Aircraft

825
1937 1947 1951

1903

1931

1931

1932

Wright Flyer Twist 1952

Pterodactyl IV Sweep 1964

MAK-10 Span 1964

IS-1 Bi-to monoplane 1966

LIG-7 Chord 1967

MAK-123 Span 1967

X5 Sweep 1969

XF10F Sweep 1970

F 111 Sweep 1972

XB 70 Span bending 1974

Su 17 IG Sweep 1974

MIG 23 Sweep 1979

SU 24 Sweep 1981

Tu 22 M Sweep 1985

F 14 Sweep 1993

FS 29 Span 1994

B1 Sweep 2001

Tornado Sweep 2002

AD 1 Obliquing 2003

Tu 160 Sweep 2004

AFTI/F 111 M.A.W. 2005

FLYRT Span 2006

MOTHRA Camber 2006

AAL Pitch 2007

F/A 18 A.A.W. 2007

Virginia Tech Span 2007

Univ. of Florida Twist 2008

Univ. of Florida Gull 2010

MFX 1 Sweep & Span

Univ. of Florida Sweep

Virginia Tech Camber

Univ. of Florida Folding

MFX 2 Sweep & span

Delft Univ. Sweep

Virignia tech Camber

Figure 1. Timeline of fixed wing aircraft implementing morphing technologies.

sliding aerodynamic surfaces, together with load-transfer attachments between the skin and the skeleton. This requires a distributed array of actuators, mechanisms, and materials that slide relative to each other or skin materials that stretch. Inman (2001) presented the actuator requirements for controlled morphing air vehicles. Mechanism design requirements include the range of motion, concerns about binding and friction, the effects of wing structural deformability under load, and the control of the actuator stroke under load. Actuator performance power and actuator force capability are essential to design success. The size, weight, and volume of the actuators are important metrics, as are the range of motion, bandwidth, and fail-safe behavior. Locking is important when the wing is under load since, without locking features, the actuators must withstand full load

unless the actuator works in parallel with a structural element. Morphing designs may also benefit from geometrically flexible structures if the aeroelastic energy from the airstream can be used to activate the shape changing and tabs can maintain the shape using aeroelastic control. The wealth of new technologies available to the wing designer provides intriguing design possibilities. Internal control and the strategy used to move from one wing form to another is also an important design goal. This involves sensor selection, braking, and locking and the integration of sensors, actuators, and the associated software. The speed at which morphing shape change occurs is also important. While slow changes may be sufficient to alter performance for some missions, rapid changes may significantly increase aircraft maneuverability.

826

S. BARBARINO ET AL.

Rotary-wing aircrafts have challenged aeronautical engineers with a plethora of issues to obtain stable flight. A major component of these issues is the complex flow field that a rotor blade is exposed to. Even in hover, each section of the rotor blade has different oncoming flow velocities, and engineers have designed the blade with a pre-built twist to compensate for them. However, the optimum amount of this pre-twist varies with the flight condition, and hence classical rotor blade designs are a compromise. Aside from the performance issues, the rotor blades emanate a significant amount of vibration and noise, caused by factors including the increased asymmetrical loading of the rotor disc, and blade tip shocks that occur with increasing forward speed. Furthermore, the noise emitted by the rotor blades is compounded by the interaction of the blades with the tip vortices of the proceeding blades (bladevortex interaction (BVI) noise) (Leishman, 2006). These problems, while not necessarily imperative to performance improvements, have to be solved in order to obtain a smooth ride. This article focuses on structural, shape-changing morphing concepts for both fixed and rotary wings, with particular reference to active systems (system characterized by an actuator). Thus, applications excluded from this study are those dealing with synthetic jets, vibration issues, and flight control, or belonging to other fields such as naval, automotive, or space. Inflatable solutions have been not considered. The wing morphing concepts have been classified into three major types, as shown in Figure 2: planform alteration (involving span, chord, and sweep changes), out-ofplane transformation (twist, dihedral/gull, and spanwise bending), and airfoil adjustment (camber and thickness). For each geometrical parameter, a table summarizes the studies available in literature on that specific topic, together with some useful information to give the reader an immediate picture of the developed work. However, some studies and past projects (for example, the DARPA (2011) Morphing Adaptive Structure (MAS) program) combine several morphing variables and are difficult to categorize. These works have been referenced in the section which best suits the principal affected parameter.

In morphing applications, where large shape changes are expected, the design of a suitable skin is a huge challenge and a key issue. The skin has to withstand the aerodynamic pressure loads, while being sufficiently compliant for the underlying morphing structure. This topic will be discussed only briefly within this article. BENEFITS OF MORPHING The current use of multiple aerodynamic devices (such as flaps and slats) represents a simplification of the general idea behind morphing. Traditional control systems (with fixed geometry and/or location) give high aerodynamic performance over a fixed range and for a limited set of flight conditions. Outside of this range, these traditional systems can be neutral or negatively influence the aerodynamics and hence often give lower efficiency. Conventional hinged mechanisms are effective in controlling the airflow, but they are not efficient, as the hinges and other junctions usually create discontinuities in the surface, resulting in unwanted fluid dynamic phenomena. Since 1920, airplanes have used devices that can increase the lift during landing and takeoff (Renken, 1985). Increases in aircraft weight and cruise speed, and increases in the wing structural stiffness to avoid multiple aero-elastic phenomena (divergence, flutter, etc.) have led to the use of discrete control surfaces such as ailerons and flaps in place of wing twist. Only since the late 1970s have researchers seriously revisited concepts of variable wing shape. Most of this research was based on two concepts, namely the active control of the curvature along the wingspan and the implementation of flexible wings, able to exploit the aero-elastic forces to obtain the desired deformations (wing shapes). Many studies have targeted the variation of the wing shape (Weiss, 2003; Sofla et al., 2010) to achieve several objectives, such as the control of shock waves during transonic flight conditions, or the control of turbulence, the wake (laminar flow separation), vortices, active control of flutter, etc. (Stanewsky, 2001). Civil and military aircrafts are designed to have optimal aerodynamic characteristics (maximum lift/drag ratio) at one point and fuel condition in the entire flight envelope. However, the fuel loading and distribution change continuously throughout the flight, and the aircraft may often have to fly at non-optimal flight conditions due to air traffic control restrictions. The consequent sub-optimal performance has more significance for commercial aircraft as they are more flexible than military aircraft and also fuel efficiency has far greater importance as a performance measure. There is also much recent interest in high-altitude long endurance (HALE) aircrafts that are designed to fly for several days. HALE aircrafts have a larger proportion of fuel weight than other aircrafts, and hence the resulting

Shape morphing wing

Planform

Sweep

Span

Chord

Out-of-plane

Twist

Dihedral/ Gull

Spanwise bending

Airfoil

Camber

Thickness

Figure 2. Organization of the review.

Morphing Aircraft

827

changes in aeroelastic shape throughout the flight can be substantial. Fuel efficiency is becoming increasingly important for civil and HALE aircrafts, and any approach that enables better aerodynamic performance throughout a flight needs to be investigated and developed. Using an adaptive wing, whose geometry varies according to changing external aerodynamic loads, the airflow in each part of the aircraft mission profile may be optimized, resulting in an increase of aerodynamic performance during cruise (Szodruch, 1985; Smith and Nelson, 1990; Siclari et al., 1996; Martins and Catalano, 1998) and maneuvers (Thornton, 1993). In the literature, the change in wing shape is also referred to as the mission adaptive wing. Even a 1% reduction in airfoil drag would save the US fleet of wide-body transport aircraft $140 million/year, at a fuel cost of $0.70/gal (source National Aeronautics and Space Administration (NASA) Dryden Studies). For a medium-range transport aircraft with an adaptive wing, the projected fuel saving should be about 35%, depending on mission distance. Large shape change concepts usually have associated design penalties such as added weight or complexity and, without these penalties, morphing would always make sense. For most applications, there is a crossover point where fuel penalties for not morphing begin to exceed the morphing weight penalty (Bowman et al., 2007). However, when overall system performance and mission requirements are assessed, large shape change concepts can be a viable approach for some missions, particularly those that combine several requirements in terms of speed, altitude, or take-off and landing. The adaptive wing would allow a given aircraft to perform multiple missions and enable a single aircraft with multirole capabilities, radically expanding its flight envelope. From a military perspective, a single morphing aircraft could perform different roles within a given mission that otherwise would require different vehicles. Jha and Kudva (2004) studied how changing wing parameters affect the performance of an aircraft, and demonstrated that an optimal design requires large geometric changes to satisfy a multi-role mission. Therefore, choosing between high efficiency or high maneuverability at the design stage would not be necessary. Aerodynamic optimization for a single flight condition would not be necessary. Compared to conventional aircraft, morphing aircrafts become more competitive as more mission tasks or roles are added to their requirements. The impact that a morphing wing has on aircraft performance for a range of flight conditions may be presented graphically using a radar (or spider) plot. Figure 3 shows a spider plot for 11 flight conditions, where each axis of the plot represents the performance metric associated with each flight condition. The outer radius of the spider plot indicates the best possible

performance. For a test case based on a BQM-34 Firebee unmanned target drone aircraft (Joshi et al., 2004), the performance of the baseline aircraft can be compared to various wing morphing strategies on the spider plot. Planform morphing (a wing capable of telescoping, chord extension, and variable sweep) significantly improves the aircraft performance over that provided by morphing the airfoil alone or by the baseline aircraft. Many researchers follow a multidisciplinary optimization and systems approach to morphing solutions (Samareh et al., 2007), focusing on the integration of computational fluid dynamics (CFD) and computational structural dynamics models for geometric optimization. Morphing wing airfoil and configuration geometry, design methodology, effectors, flight control, aero-elasticity, and stability may also be considered (Rodriguez (2007) gave an overview). Bowman et al. (2007) demonstrated that for a hunter-killer mission, that is mainly cruise- and loiter-dominated, morphing does make sense (even with a 10% empty weight penalty), especially for vehicles under 20,000 lbs. For more difficult missions, for example, increasing the search radius, this break point in vehicle weight would increase. Roth et al. (2002) showed that morphing could have a large impact on fleet size for a Coast Guard patrol mission. The key to requirements for this mission was a high-altitude cruise out to station with a fast response time, and then a slow, low-altitude patrol. Smith et al. (2007) investigated the benefits of morphing wing technology for fighter aircraft systems (with both turbojet and turbofan propulsions), showing that the optimized mission-integrated designs yield significant fuel savings over fixed-wing aircraft. Mission segment analysis shows that most fuel savings due to morphing are obtained in the least-constrained (subsonic) flight segments. Wittmann et al. (2009) gave a methodology to assess morphing strategies and technologies for a wide range of applications, including civil and military, using a variety of figures of merit to evaluate achievable benefits. For instance, a single degree-of-freedom variable-chord morphing concept evaluated for a high-speed scenario yielded 79% morphing efficiency for the maximum speed as a figure of merit. An efficient transport scenario saw a 23% improvement in lift-to-drag ratio when the morphing parameters were optimized simultaneously. For a high-lift scenario, a 74% increase in lift coefficient could be achieved by maximizing wing area and camber. From a flight control perspective, the pitching moment is strongly influenced by wing sweep (and can lead to a reduction of the trim surfaces), while roll and pitch control can be efficiently modified by variable span (and also variable twist). Adaptive wings can provide a significant increase in performance. However, the realization of such an

828

S. BARBARINO ET AL.

Firebee Airfoil Takeoff: SL S-Turn: 60k Climb: SL goemetry

I-Turn: SL

Climb: 30k

Loiter: 60k

Cruise: SL

Dash: 30k

Cruise: 30k

Accel: 30k

Cruise: 60k

Figure 3. Spider plot comparing predicted performance of the baseline Firebee, a morphing airfoil Firebee and a morphing planform Firebee (Joshi et al., 2004, reprinted with permission of the American Institute of Aeronautics and Astronautics).

aircraft poses new design challenges. In traditional design approaches, the airfoil profile and wing shape are dictated by the aerodynamics at fixed flight conditions and the structure is lightweight and extremely rigid. This approach must be abandoned, and innovative structures designed that are inherently deformable and continuously adaptable, in real time and during flight, according to the operational conditions and requirements imposed by the mission. Singh et al. (2009) gave a methodology to create innovative products, labeled as transformers, with a broader functional repertoire than traditional designs, and able to transform into different configurations. They highlighted the lack of a systematic methodology to create such products and identified analogies in nature, patents, and products. They also hypothesized the existence of such products in different environments and situations. Increasing wing structural weight is a serious problem. Skillen and Crossley (2006, 2008b) showed in their numerical study that mechanisms can account for a substantial portion of the weight and are a strong function of the wing geometry. Moreover, the opportunity to realize a morphing wing requires the availability of

materials and implementation solutions that guarantee in all circumstances the necessary deformation of the structure while maintaining structural integrity and load-bearing capability. Smart materials may be the solution to realize distributed actuators within the wing structure capable of providing shape modifications. The current challenges of morphing vehicle design are the additional weight and complexity, the power consumption of the required distributed actuation concepts, and the development of structural mechanization concepts covered by flexible skins (Reich and Sanders, 2007). Additionally, there is a strong need to understand the scalability of morphing wing concepts to achieve sufficient structural stiffness, robust aero-elastic designs, and an adequate flight control law to handle the changing aerodynamic and inertia characteristics of morphing vehicles (Moorhouse et al., 2006). Flight control represents another big challenge for morphing aircraft, as additional terms appear in the governing equations and may require complex control systems. Seigler et al. (2007) studied the modeling and flight control of vehicles with large-scale

Morphing Aircraft

829

planform changes. The equations of atmospheric flight were derived in a general form, methods to integrate the aerodynamic forces examined, and various approaches and methods of flight control distinguished. The absence of sharp edges and deflected surfaces on morphing aircraft also has the potential to reduce the radar signature and visibility of the vehicle, thus enhancing its stealth properties. Gern et al. (2005) highlighted that redundancy generally requires a conventional flap to be connected to at least one or two additional actuators to provide sufficient flight control in case of airframe damage or actuator failure. This may dramatically increase the power installed in a vehicle; for instance, the B-2 stealth bomber has 24 actuators to drive 11 primary control surfaces. In contrast, morphing wings usually have heavily distributed actuation that provides sufficient robustness and redundancy to account for actuator failures with only a negligible effect on weight. Current research on rotary wings focuses on improvements in terms of increased speed, payload, and maneuverability, along with reductions in costs, vibrations, and noise. The DARPA Mission Adaptive Rotor (MAR) initiative (Warwick, 2009) plan to fly an adaptive rotor by 2018. Boeing, Sikorsky, and the Bell Boeing tiltrotor team have received 16-month Phase 1 contracts to assess a wide range of adaptive rotor technologies and develop designs for both a completely new rotor system and a retro-fit demonstrator rotor. The goal of the MAR project is a rotor that can change its configuration before a mission and in flight, between mission segments and with every revolution. DARPA has identified a wide range of potential approaches to reconfigure the rotor in flight, including varying the blade diameter, sweep, and chord; morphing tip shapes and variable-camber airfoils; varying blade twist, anhedral/dihedral, tip speed, stiffness, and damping. DARPA is looking for technologies that can increase rotorcraft payload by 30% and range by 40%, while reducing the acoustic detection range by 50% and vibration by 90% over fixed geometry rotors. Some studies demonstrate the aerodynamic and acoustic benefits of an active twist rotor (Bailly and Delrieux, 2009; Boyd and Douglas, 2009; Zhang and Hoffmann, 2009), rotor power reduction using trailing-edge flaps (Leon and Gandhi, 2009) and envelope expansion using extendable chord sections (Leon et al., 2009). This article only includes morphing applications with performance objectives. MORPHING SKINS Most of the morphing technologies or concepts assume the existence of an appropriate flexible skin. The design of flexible skins is challenging and has many conflicting requirements. The skin must be soft enough to allow shape changes but at the same time it

must be stiff enough to withstand the aerodynamic loads and maintain the required shape/profile. This requires thorough trade-off design studies between the requirements. In addition, the type of flexible skin required depends on the loading scenario and the desired change in shape (one-dimensional or multi-dimensional). Therefore, the design of flexible skins depends on the specific application. Thill et al. (2008) performed a comprehensive review of flexible skins and considered various novel material systems concepts and technologies. Only papers published since this review are briefly described here. Thill et al. (2010) investigated the use of a composite corrugated structure as morphing skin panels in the trailingedge region to vary the camber and chord of an airfoil. Wereley and Gandhi (2010) examined extensively the challenges and requirements of flexible skins and listed the state-of-the-art solutions. Skin concepts including elastomeric matrix composites have been investigated for large area changes (Peel et al., 2009; Bubert et al., 2010; Murray et al., 2010, Olympio et al., 2010). Furthermore, the use of morphing core sandwich structures covered by compliant face sheet has been investigated for both low- and for high-strain applications depending on the cellular arrangement and the material of the core (Joo et al., 2009; Bubert et al., 2010, Olympio and Gandhi, 2010a, b). In addition, McKnight et al. (2010) studied the use of segmented reinforcement (discrete) to create a variable-stiffness material that can be adopted as a flexible skin.

WING PLANFORM MORPHING Wing planform is mainly affected by three parameters (individually or in combination): span, chord, and sweep. Both span and sweep can affect the wing aspect ratio, a parameter that modifies the lift-to-drag ratio. Thus, an increase in wing aspect ratio would result in an increase in both range and endurance (Anderson, 2000). From an aerodynamic perspective, the change in aspect ratio produces differences in lift curve slope and forces due to the change in the wing area. From a dynamics perspective, the inertia of the aircraft also changes. Span and sweep morphing have been investigated for military applications, particularly for UAVs, which must loiter during surveillance and rapidly switch into high-speed dash mode to move reconnaissance area or to attack a target. Chord morphing has mainly been applied to helicopter rotor blades so far. The objective of the DARPA MAS program was to design and built active, variable-geometry wing structures with the ability to change wing shape and wing area substantially. All three MAS contractors (Lockheed Martin, Hypercomp/ NextGen, and Raytheon Missile Systems) conducted a functional analysis that concluded that changing wing

830

S. BARBARINO ET AL.

planform area and wingspan were the primary enablers of a new class of morphing air vehicles. Tables 13 summarize the literature for each geometrical parameter. However, due to the intrinsically coupled nature of many applications of planform change, some papers appear in more than one table. Variable Span for Fixed Wing Aircraft Wings with large spans have good range and fuel efficiency, but lack maneuverability and have relatively low cruise speeds. By contrast, aircrafts with low-aspect ratio wings are faster and highly maneuverable, but show poor aerodynamic efficiency (McCormik, 1995). A variable-span wing can potentially integrate into a single aircraft the advantages of both designs, making this emerging technology especially attractive for military UAVs. Increasing the wingspan increases the aspect ratio and wing area and decreases the span-wise lift distribution for the same lift. Thus, the drag of the wing decreases, and consequently, the range of the vehicle increases. Unfortunately, the wing-root bending moment can increase considerably due to the longer span. Thus, both the aerodynamic and the aeroelastic characteristics should be investigated in the design of variable-span morphing wings. Most span morphing concepts are based on a telescopic mechanism, following the ideas of the Russian expatriate Ivan Makhonine, where the wing outer panel telescoped inside the inner panel to enable span and wing area changes. The MAK-10 was the first design with a telescopic wing and it first flew in 1931. The mechanism was powered pneumatically and enabled span increased up to 62% (from 13 to 21 m) and area increased up to 57% (from 21 to 33 m2) (Weisshaar, 2006). The US Department of Defense has designed numerous vehicle configurations in the past 20 years to meet various mission objectives. Bovais and Davidson (1994) from the Naval Research Laboratory built an experimental non-recoverable ship-launched expendable radar decoy named FLYRT (Flying Radar Target) that was first flown in September 1993. FLYRT was launched with rigid folded wings and tail surfaces from an MK 36 launcher using a solid-propellant rocket motor which burned for about 1.6 s. The fully expanded rigid wing had a span of 2.4 m and weighed 60 kg. Immediately after launch, the tail fin unfolded mechanically to control the vehicle during ascent. A total of 13 drones were built before the program ended, and the decoy successfully demonstrated the defense of a variety of ships against simulated radar threats. Gevers Aircraft Inc. (1997) investigated a telescopic wing, capable of variable span, for use on a six-seat amphibious aircraft. The wing was composed of a fixed center section and two extendable outer sections,

using an overlapping extension spar system. The center section was a high-speed wing (low drag and strong) and the completely retractable high-lift section moved in the span-wise direction. The morphing unmanned aerial attack vehicle, developed by AeroVisions Inc. (2011; Website) within the Morphing Aircraft Structures program funded by DARPA, consisted of several sliding segments. The wingspan was inversely proportional to the cruise speed, and allowed for several operating conditions from loitering to fast cruise to high-speed attack. Bae et al. (2005) performed both static aerodynamic and aeroelastic studies on the wing of a long-range cruise missile and highlighted some of the benefits and challenges associated with the design of a morphing wing capable of span change. The total drag of the morphing wing decreased by approximately 25%, and the range increased by approximately 30%. The aeroelastic analysis showed that the flexibility of the morphing wing structure increased as the wingspan increased. At a given flight condition, the deformation from the aerodynamic loads was much larger than that of the conventional wing. Static aeroelastic considerations showed that a variable-span wing requires increased bending stiffness because the bending deformation is more significant than twist. The Air Force Research Laboratory (2011; AFRL) Vehicle Research Section (Website) developed an unmanned planform called ALICE (Air Launched Integrated Counter-measure, Expendable) that can be air launched from a tactical aircraft at speeds up to Mach 0.8 and altitudes up to 45,000 ft. After launch, ALICE glides using tail control surfaces until it reaches a speed of 250 knots. The rigid wing and propeller then deploy and the heavy fuel engine starts. ALICE will cruise approximately 200 nautical miles in 1 h before the outer rigid wing panels deploy for loiter. Research advances included the development of the polymorphic wing, a JP-8 fueled rotary engine, a high-efficiency starter/generator, a folding variable-pitch propeller, and an advanced EW payload. Arrison et al. (2003) modified a Delta Vortex RC aircraft by adding telescopic wings. The new vehicle, renamed BetaMax, could increase wingspan by 10 in. (over a 43.5 in. basic span). A range increase of 19% over the conventional aircraft was predicted, allowing for the increased weight. The RC vehicle was successfully flown, and it highlighted a change in static stability between the retracted and extended case of nearly 5%. Neal et al. (2004) designed and demonstrated a variableplanform aircraft capable of large wing planform changes, both in sweep and span, using a telescopic pneumatic actuator. The aspect ratio could change up to 131% through combined span and sweep, while wing area could change by 31%. Wind tunnel results showed that only three planform geometries were required to

Table 1. Planform morphing: Span.


Vehicle Investigation Experimental Category Rotary wing    8      Rotary wing Helicopter Rigid Civil tiltrotor Sliding     8 8 Size Actuation Skin Purpose Numerical Prot. WT 8   8 8 8 FTB 8 8

References

Geometrical parameters

Span

Span

Strap around a drum (main rotor) Jackscrew (main rotor)

Performance (disc loading, rotor tip speed) Performance (drag at high forward speeds) Performance (cruise and hover efficiency)

Span Fixed wing Rotary wing Rotary wing Civil tiltrotor Civil tiltrotor UAV Rigid

Rotary wing

Civil tiltrotor

8  8 8

Martin and Hall (1969) Fradenburgh (1973) Frint (1977) Studebaker and Matuska (1993) Bovais and Davidson (1994) Davis et al. (1995)

Span

Span

Brender et al. (1997)

Span

Wang et al. (1999)

Span

Rotary wing

Civil tiltrotor

Jackscrew (electric)

Rigid

Morphing Aircraft

Span Span

Fixed wing Fixed wing

UAV UAV

Rack and pinion Pneumatic

Sliding Sliding

Performance (cruise and hover efficiency) Performance (disc loading and rotor tip speed) Performance (disc loading and rotor tip speed) Performance (L/D) Performance (L/D)

 

8 

 8

Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing UAV UAV UAV

UAV

Electromechanical

Sliding Sliding Sliding Sliding

   Performance (L/D, range) Rigid Performance (L/D)  

   8 

8   8 

8 8 8 8 8

Arrison et al. (2003) Blondeau et al. (2003) Blondeau and Pines (2007) Sullivan and Watkins (2003) Neal et al. (2004, 2006) Alemayehu et al. (2004, 2005) Bae et al. (2005) Pneumatic and electromechanical DC motor and servos Cruise missile MAV DC motor Pneumatic Flexible

Obliquing and Span Span, sweep, and twist Span and sweep

Performance (L/D) Control (roll) Performance (L/D)

Span

Span and camber

Span and sweep Span and airfoil shape

Performance (CD)

 

 

8 8

8 8

Heryawan et al. (2005) Joo et al. (2006) Vale et al. (2006, 2007) Gamboa et al. (2007) Bharti et al. (2007) Fixed wing UAV

Sweep and span

DC motor

8 (continued)

831

832

Table 1. Continued.
Vehicle Investigation Experimental Category Fixed wing UAV Rotary mechanical system Electric motor Servos UAV UAV SMP hinge UAV UAV UAV Pneumatic Pneumatic Flexible Sliding Variable RPM (main rotor) Sliding Flexible Performance (L/D) Hybrid (rigid and flexible) Performance (L/D) Size Actuation Skin Purpose Numerical   8  Performance (L/D) Performance (L/D) Performance (CD) Performance (L/D) Performance (L/D) Hybrid (rigid and flexible)       Prot.       8    WT    8 8 8 8 8 8 8 FTB 8

References

Geometrical parameters

Span and sweep

Sweep and span Fixed wing Rotary wing Fixed wing Fixed wing Fixed Wing Fixed wing Fixed wing Fixed wing

Fixed wing

UAV

 8 8 8 8 8 8 8 8

Bye et al. (2007) Ivanco et al. (2007) Love et al. (2007) Andersen et al. (2007) Flanagan et al. (2007) Han et al. (2007) Performance (L/D) Control (roll) Performance (L/D)

Span

S. BARBARINO ET AL.

Prabhakar et al. (2007) Supekar (2007) Yu et al. (2007) Gamboa et al. (2009)

Span

and gull and sweep and shape

Leite et al. (2009) Scott et al. (2009)

Span Span Span airfoil Span Span

Grady (2010)

Span and sweep

Morphing Aircraft Table 2. Planform morphing: Chord.


Vehicle Geometrical Year parameters Category Chord Chord Fixed wing Rotary wing Rotary wing Rotary wing Investigation

833

Experimental Authors Perkins et al. (2004) Reed et al. (2005) Leon et al. (2009) Leon and Gandhi (2009) Barbarino et al. (2010b) Khoshlahjeh et al. (2010, 2011) Size UAV Actuation DC motor Skin Purpose Numerical Prot. WT FTB    8 8 8 8 8

Flexible Performance (CL)  Rigid Performance (flight envelope)   

Helicopter DC motor

Chord Chord

Helicopter Variable RPM Flexible Performance (main rotor) (flight envelope) Helicopter Performance (flight envelope)

8 8

8 8

Table 3. Planform morphing: Sweep.


Vehicle Geometrical parameters Sweep Span and sweep, twist Span and sweep Sweep Sweep Investigation Experimental Category Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Size UAV UAV UAV MAV MAV Servos Rigid Actuation Pneumatic Pneumatic and electromechanical DC motor and servos Skin Rigid Sliding Sliding Performance (L/D) Control (turn radius and crosswind rejection) Purpose Performance (L/D) Performance (L/D) Numerical      Prot.    8  WT    8 8 FTB 8 8 8 8 8

References Marmier and Wereley (2003) Neal et al. (2004, 2006) Alemayehu et al. (2004, 2005) Hall et al. (2005) Grant et al. (2006)

Joo et al. (2006) Mattioni et al. (2006) Bharti et al. (2007) Bye et al. (2007) Ivanco et al. (2007) Love et al. (2007) Andersen et al. (2007) Flanagan et al. (2007) Yu et al. (2007) Grady (2010) Sofla et al. (2010)

Span and sweep Sweep Sweep and span Span and sweep Sweep and span

Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing

Pneumatic

 

    

8 8 8  

8 8 8 8

UAV UAV

DC motor Rotary mechanical system Electric motor Hybrid (rigid and flexible) Flexible Performance (L/D) Performance (L/D)

  

UAV

Span and sweep Span and sweep Sweep

Fixed wing Fixed wing Fixed wing

SMP hinge Pneumatic UAV SMA

Performance (L/D)

 

  

8 8 8

8 8 8

Performance (power consumption)

834

S. BARBARINO ET AL.

maintain minimum drag over a range of possible lift coefficients. Neal et al. (2006) redesigned the vehicle to implement a variable-geometry tail and increased the strength of the structure and mechanisms. Blondeau et al. (2003) designed and fabricated a threesegmented telescopic wing for a UAV. Hollow fiberglass shells were used to preserve the span-wise airfoil geometry and insure compact storage and deployment of the telescopic wing. To reduce the weight, they replaced the wing spars with inflatable actuators that could support the aerodynamic loads on the wing. Their telescopic spar design consisted of three concentric circular aluminum tubes of decreasing diameter and increasing length, connected by ceramic linear bearings, and deployed and retracted using input pressure. The wing could undergo a 114% change in the aspect ratio, while supporting aerodynamic loads. Wind tunnel test results showed that the wing suffered from parasitic drag created by the seams of the wing sections. Fully deployed the telescopic wing achieved lift-to-drag ratios as high as 9 to 10, which is approximately 25% lower than its rigid fixed wing counterpart. In a further development, Blondeau and Pines (2007) adopted two identical telescopic spars instead of one, mechanically coupled by the ribs, to prevent wing twist and fluttering. The new prototype could undergo a 230% change in aspect ratio, and seam heights were reduced giving less parasitic drag. In its fully deployed condition, the telescopic wing could achieve lift-to-drag ratios as high as 16, which was similar to its solid foam-core wing counterpart. Supekar (2007) evaluated the structural and aerodynamic performance of a two-segment telescopic wing for a UAV that could also undergo variable dihedral of the outer wing. Although a crude prototype was manufactured, no successful actuation of the mechanism was reported due to the fabrication problems. Sullivan and Watkins (2003) investigated an oblique wing aircraft (an aircraft capable of rotating its wing around a central pivot) that was capable of symmetric and asymmetric span extensions. With only symmetric span extension, the aspect ratio could change between 3.3 and 4.7, leading to an increase in stall and cruise speed (approximately 30%) and an increase in take-off ground roll (100% increase). The Virginia Tech Morphing Wing Team (Alemayehu et al., 2005) developed a morphing wing with three different morphing characteristics: change in span, sweep angle, or chord length. Heryawan et al. (2005) designed an expandable morphing wing for a micro-aerial vehicle (MAV) based on a typical bird wing. The wing was divided into two rigid parts (made of carbon composite and balsa at the leading edge, and carbon fiber composite mimicking wing feathers on the remaining part); the outer wing was driven by a small direct current (DC) motor through reinforced composite linkages. The wing was able to

change aspect ratio from 4.7 to 8.5 in 2 s. Additionally, two lightweight piezo-composite actuators (LIPCAs) were attached under the inner wing to modify camber. Wind tunnel tests carried out for a 7.5 m/s wind speed (Re 30,000) and at angles of attack between 0 and 8 showed that the total drag decreased significantly (more than 10%) during wing expansion while the lift more than trebled. The maximum value of lift-to-drag ratio occurred at 2 angle of attack for the low-aspect ratio configuration, in contrast to 0 for the high-aspect ratio configuration. Finally, the actuation of the two LIPCAs produced an additional 16% lift. Han et al. (2007) performed wind tunnel tests on a model with telescopic wings actuated by two separate servomotors to study the effect of variable aspect ratio on wing-in-ground effect vehicles operating inside a channel. Changing the aspect ratio from 3.2 to 3.5 improved the lift-to-drag ratio more than the effect of both the ground and the sidewalls. Span changing also had a bigger advantage with walls present (up to 54.7% lift-to-drag increase compared to the original wing). Wing tip extension does not control rolling moment efficiently on its own, but the influence of the ground or sidewall effects generates a positive rolling moment due to the high-pressure air trapped between the lower surface of the wing and the ground (or sidewall). An alternative approach to changing wingspan uses a scissor-like mechanism for the wingbox. Joo et al. (2006) studied the optimal location of a distributed network of actuators within such a mechanism. Their reconfigurable wingbox was constructed of four-bar mechanisms with rigid links, and an experiment was conducted on a single cell linkage using a pneumatic actuator. An optimization analysis was performed to select the optimal actuator placement. Springs were used in their design to account for a stretchable skin, but a parametric study of the compliance was not performed. Johnson et al. (2009) further developed this work by exploring the effect of the optimal actuator placement and position on energy efficiency in morphing wings. Bharti et al. (2007) explored a scissor-mechanism to alter the aircraft span and sweep with a design based on the TSCh wing (NextGen Aeronautics Inc.). A scale prototype mechanism was constructed that achieved a 55% span change. Vale et al. (2006, 2007) and Gamboa et al. (2007) designed a wing section capable of independent changes of span and airfoil chord and shape by the use of extendable ribs and spars. They compared the performance achievable with such a morphing wing, in terms of minimum drag, to a traditional fixed wing at different flight speeds (1550 m/s). An aero-structural analysis was performed, considering a mechanism that could expand in the span-wise direction, keeping the ribs evenly distributed, and increasing the chord using the ribs. The skin was assumed to be rubber, wrapped as a sleeve around the wing internal mechanism and structure, with

Morphing Aircraft

835

some pre-tension. However, the aerodynamic loads acting on the skin, and airfoil thickness and area, reduce between consecutive ribs (where it is unsupported), leading to a non-optimal shape that also affects parasitic drag (for instance, at 30 m/s, the reduction in drag is 22.5% when compared to the non-morphing wing). The final design was also heavy due to the servomotors, transmission components, and other equipments; moreover, the torque actuation requirements for the chord expansion mechanism were prohibitive. Gamboa et al. (2009) proposed new structural designs for chord and span extension, but no prototype results are available. Leite et al. (2009) further investigated the concept of a telescopic wing with airfoil shape change capability, although their prototype had only span change. The airfoil change was assumed to be between two different and specific airfoils (Eppler 434 and NACA 0012). The coupled aero-structural analysis studied the optimal combination of morphing parameters within a typical mission profile for a proposed remote piloted vehicle, producing an optimum morphing wing polar curve that outperformed the original vehicle in terms of extra take-off weight and lower drag at all speeds. Scott et al. (2009) developed a novel low-stored volume wing for a HALE aircraft using a hybrid design incorporating a telescoping and load-bearing spar within an inflatable wing. The primary benefits include significantly lower required inflation pressures and lower overall system weight. The overall goal was to develop a low-stored volume wing design with a projected vehicle lift-to-drag ratio greater than 27 at altitudes ranging from 60,000 to 75,000 ft. The sequence from initial ballistic delivery to complete vehicle deployment is as follows: the aircraft is separated from a rocket in the stowed position either by mechanical means or a parachute; the tail deploys and the rigid wing rotates 90 to provide a stable craft capable of weathering the deployment dynamics; the span extensions deploy, giving the vehicle four times the span of the stowed rigid wing. The design of highaspect-ratio wings is driven by the ability to sustain substantial root bending moments: to bear these bending moments with a purely inflatable wing would require very high (50100 psi) pressurization causing problems specifying the skin fabric and in vehicle certification. The hybrid approach utilizes ribfoils and tension wires to support the skin, thereby reducing the chamber pressure to the minimum required to maintain the airfoil profile. The telospar (hybrid structural actuator) serves as the primary load path as the wing section is deployed, and this concept has proven viable in a series of deployment tests, in ambient and vacuum conditions. Systemlevel studies concluded that this deployed span design would reduce pressurization requirements by almost an order of magnitude, and reduce wing weight by 2535% (compared to a purely inflatable vectran-baffled wing).

The most dramatic morphing wing involving span change that has been realized as a wind tunnel prototype is the Agile Hunter by Lockheed Martin (Bye and McClure, 2007; Ivanco et al., 2007; Love et al., 2007). Funded by DARPA within the MAS program, the prototype (also called a Z-wing) was based on a military UAV capable of folding the inner sections of the wing near to the fuselage to reduce the surface area and drag during transonic flight at low altitude. The major challenge was the realization of suitable hinges that connect the two wing portions; the hinges have to sustain the aerodynamic loads but offer a smooth, continuous aerodynamic surface. Several materials were considered, including silicone-based and shape memory polymer (SMP) skins. Wind tunnel tests at Mach 0.6 showed a morphing capability from 0 to 130 over 65 s with a controllable, reliable, and precise actuation. Asymmetrical span morphing can be used for roll control. Henry and Pines (2007) extended the standard aircraft dynamics model to include the additional terms (as perturbations) due to morphing and demonstrated that asymmetrical span morphing was effective for roll control. The total damping in the system increased when the span extension rate was positive (span increase) due to the conservation of angular momentum. Span extension induces a roll damping moment that is greater than that due to aileron deflection. Seigler et al. (2004) also investigated asymmetrical span extension for increased maneuverability of bankto-turn cruise missiles. By formulating a full non-linear model of the missile, due to the shift of the missiles center of mass and the dependence of the rolling moment on the angle of attack, they showed that the control authority can be significantly larger when compared to conventional tail surface control. Improved maneuverability, however, is highly dependent on the angle of attack, linear actuation speed, and extension length. Moreover, as the mass of the extending wings becomes large relative to the missile body, the rigid body dynamics can become increasingly complex. The paper presented a non-linear control law to control the roll, angle of attack, and sideslip angles in accordance with bank-to-turn guidance. The control method proved to be adept in tracking-commanded inputs while effectively eliminating sideslip. Variable Span for Rotary-Wing Aircraft The variable-diameter rotor (VDTR) is not a new concept. In the late 1960s, the telescoping rotor aircraft (TRAC) rotor system was designed and tested for use in stowed rotor and compound helicopter applications. For stowed rotors, the design alleviated dynamic and strength problems associated with stopping a rotor during flight, while for compound helicopters, the design reduced drag at high forward speeds.

836

S. BARBARINO ET AL.

The TRAC rotor used a jackscrew mechanism to slide an outer blade section over an inner aerodynamically shaped tube. Wind tunnel tests and actuator mechanism cycle tests demonstrated the performance benefits and the feasibility of the concept (Fradenburgh, 1973; Frint, 1977). Prabhakar et al. (2007) used the inherent centrifugal force for actuation, reducing power requirements and mechanical complexity. The span extension increases with rotor speed and the amount is calibrated during the design phase. A scaled prototype was built and tested. VDTRs could solve many of the technical problems that arise in large civil tiltrotors. The rotors could change diameter in flight, so that a large diameter (helicopter size) rotor is used in hover and a smaller diameter (propeller size) rotor is used in cruise (Brender et al., 1997). The low disc loading in hover and low tip speeds in cruise made possible by the diameter change could eliminate many of the undesirable conventional tiltrotor attributes (Fradenburgh and Matuska, 1992). Low-disc loading tiltrotor designs have many of the advantages of helicopters during vertical flight including low power requirements (as much as 30% lower for the VDTR), low downwash velocities, good autorotation performance, and reduced BVI noise during descent. The advantages of the VDTR during cruise are derived from the low rotor tip speeds made possible by the small diameter. Low tip speeds reduce compressibility drag and lead to higher blade loading during cruise. Both factors contribute significantly to rotor propulsive efficiency (lower fuel consumption). Moreover, the VDTR, in contrast to conventional tiltrotors, does not require rotor speed reduction for efficient cruise flight as the necessary tip speed reduction may be obtained by reducing the rotor diameter. Other important advantages are reduced gust response due to higher blade loading and reduced internal cabin noise levels (i.e., improved passenger comfort) due to low tip speeds and large rotor tip distance from the fuselage. However, the diameter change mechanism adds weight and complexity to the rotor system. Martin and Hall (1969) describe a VDTR concept developed specifically for a tiltrotor aircraft. The outer blade sections of this design telescoped into the inner blade section and were actuated using a strap wound around a drum attached to the drive shaft. A 25-ft diameter version of this rotor design demonstrated 700 extensions and retractions during ground tests. Studebaker and Matuska (1993) performed a reducedscale wind tunnel test of a rotor designed for a 38,600 lb aircraft with a 36-passenger payload. This test used a 1/6 scale semi-span model to demonstrate the aeroelastic performance of the VDTR converting from hover to cruise. No instabilities encountered in any flight mode. The test also verified that the rotor root cut out had only a small effect on the hover figure of merit and that the

VDTR had improved gust response characteristics over conventional designs. The baseline VDTR design was later modified to incorporate NASA Short Haul (Civil Tiltrotor) guidelines. Davis et al. (1995) developed a dual-point optimizer to simultaneously optimize the rotor design for hover and cruise, based on the EI-IPIC/HERO (Evaluation of Hover-Performance using Influence Coefficients/ Helicopter Rotor Optimizer) design tool. Several designs showed significantly higher cruise efficiency when compared to conventional tiltrotor designs. The calculated figures of merit were only slightly lower for the VDTR designs compared to conventional designs and the corresponding hover power loading (thrust/hp) was about 30% lower for the VDTR due to its inherent lower disc loading. Wang et al. (1999) showed the potential of the VDTR concept to reduce the prop-rotor-whirl flutter and the flutter airspeed was increased by 54% for a hypothetical civil tiltrotor design. DARPA (Website) is currently funding the DiscRotor Compound Helicopter program, whose goal is to develop a new type of compound helicopter capable of high-efficiency hover, high-speed flight, and seamless transition between these flight states. The aircraft is equipped with an aft-swept wing, together with a midfuselage disc with extendable rotor blades, enabling the aircraft to takeoff and land like a helicopter. Transition from helicopter flight to full fixed-wing flight is achieved by fully retracting the blades within the disc. An aircraft capable of long-range high speed and vertical takeoff/ landing and hover will satisfy an ongoing military interest. Extending and retracting telescopic rotor blades are an important DiscRotor-enabling technology. Chord Morphing The chord length of a conventional aircraft wing is altered using leading-/trailing edge-flaps, usually actuated by screw systems. Many of these devices are patented and operational, and few alternatives have been considered. In fixed-wing aircraft, the presence of spars, fuel tanks, and other components presents additional structural complexity; in contrast, the blades of rotarywing aircraft are characterized by a single D-spar and honeycomb filler. This, together with the larger displacements required for morphing and aerodynamic loads, has restricted most chord morphing applications to rotary-wing aircraft. The Bakshaev LIG-7 was designed in the USSR in 1937 and was probably the first aircraft capable of chord increase (Weisshaar, 2006). The telescopic mechanism consisted of six chord-wise overlapping wing sections, which were retracted and extended using tensioned steel wire, driven manually from the cockpit. All retractable sections were completely hidden inside the fuselage when retracted with an area change of 44%.

Morphing Aircraft

837

Reed et al. (2005) used an interpenetrating rib mechanism to change the chord length by means of miniature DC motors and lead screws. They used partial rib structures that could slide through a central slotted box and alter the chord-wise position of the leading and trailing edges. The ribs were designed to support camber bending due to the aerodynamic loads; an additional flexible honeycomb could be attached to these ribs to maintain the airfoil shape. The smooth operation of the mechanism under aerodynamic loads and maintaining the chord-wise bending stiffness are significant challenges. The added weight and complexity of the design are disadvantages. A solution for the skin using SMPs was suggested with thin wires for heating. The application of smart materials to achieve chord change has received little attention. The Cornerstone Research Group experimented with dynamic modulus foam (DMF) to alter chord length (Perkins et al., 2004) and their goal was to achieve an 80% increase in lift. DMF foam is a lightweight form of SMP with similar behavior. An SMP was proposed for the skin, to accommodate large strains, and temperature activation was realized using thin nichrome wires embedded in the skin. Although the prototype wing section was able to extend the chord upon heating, it could not return to its original shape upon cooling because of the low recovery stress of shape memory foams. Leon et al. (2009) quasi-statically increased the chord through the extension of a flat plate (named the trailingedge plate or TEP) through a slit trailing edge over a section of a rotor blade. The objective was to increase the maximum speed, altitude, and gross-weight, and reduce the main rotor power near envelope boundaries. The concept, referred to as the static extended trailing edge (SETE) appeared to give better high-lift performance than trailing edge or Gurney flaps at high lift coefficients. Simulation results, based on a UH-60 helicopter, indicated that in stall-dominated conditions, increases of up to 3000 ft in the maximum altitude, 2400 lbs in the maximum gross-weight, and 26 knots in the maximum speed, and reductions of up to 33.4% in the rotor power, may be obtained. They also demonstrated the feasibility of a morphing X-truss mechanism to actuate the extendable plate. Barbarino et al. (2010b) extended this work by substituting the flat plate with a morphing cellular structure, suitably designed to accommodate large deformations and undergo cyclic actuation. The chord may be increased by almost 30%. The flexible pre-strained skin must maintain a smooth, continuous shape in the presence of aerodynamic loads, and several methods to attach the skin to the cellular honeycomb were investigated. The inherent centrifugal force was proposed for actuation, using a variable-speed main rotor, thus reducing the overall complexity. Khoshlahjeh et al. (2010, 2011) investigated the chord extension of rotor blade for a utility helicopter, adopting

the TEP concept of Leon and Gandhi (2009) deployed at 2 downward incidence between 63% and 83% of the blade span, and able to increase the airfoil chord by 20%. The aerodynamic modeling of airfoil sections with TEP was improved by adding CFD calculated deltas to the baseline aerodynamic coefficients. Initially, they performed analyses on a rigid blade with CFD data for the blade sections with chord extension. An elastic blade was considered and the nose-down pitching moment applied on the blade caused by the extended chord was accounted for. They also included the LeishmanBeddoes semi-empirical dynamic stall model in order to more accurately predict performance. Estimates were performed at moderate cruise speed and high-speed conditions for various aircraft gross weights and operating altitudes. The TEP extension allowed large increases in aircraft gross weight at high altitude, or alternatively significant power reductions. In stalldominated conditions, where the TEP extension is most effective, it significantly reduced the blade section angles of attack on the retreating side, and although drag increased over span-wise regions where the TEP was deployed, the overall reduction in drag due to the change in rotor trim reduced the power required. The increase in maximum speed was more modest. Variable-Wing Sweep The notion of a swept wing is not a novel idea. After World War II, aviation entered the jet age, and the resulting increases in flight speed represented a new tactical advantage. The only way to reduce the significant drag at high subsonic speeds was to use a low thicknessto-chord ratio and low-aspect ratio wings, until the allies found some evidence of German swept wing activity to confirm work published by Jones (1945). Before this, Adolf Busemann in 1935 and Michael Gluhareff in 1941 investigated aerodynamics phenomena at supersonic speeds (Loftin, 1985). New aircraft designs started to incorporate sweep, even though they suffered from a pitch-up moment at high angles of attack and mediocre handling qualities at low speeds. The idea of a variablesweep wing (or swing-wing) was born and offered a compromise: to combine efficient low-speed (for takeoff and landing) and high-speed (fast cruise or supersonic capability) flights. Pivoting of the wing has been the method of choice for the sweep change. Variable sweep changes the wing aspect ratio and area, affecting the aerodynamic forces; the lift curve slope tends to decrease as sweep increases, due to the change in aspect ratio. Performance is improved at high velocities, where compressibility effects are important, by delaying the drag rise at Mach numbers close to unity and also the buffet onset. Full-scale development of variable-geometry wings began in Germany during World War II with the

838

S. BARBARINO ET AL.

Messerschmitt P-1101, which had preset wing sweep (the sweep angle was set on the ground). The first swing-wing aircraft, the Bell X-5, flew in 1951 and was adapted from the P-1101, with a variable-sweep mechanism (Perry, 1966) actuated by a jackscrew assembly. This aircraft tended to uncontrollable spins at stall and the sweep mechanism itself was not very efficient. The wing was swept and translated forward simultaneously to control the position of the aerodynamic center (Kress, 1983). The Grumman XF-10-F used a similar mechanism but never entered service. Swing-wings became viable in the mid-1950s when the NASA Langley Research Center developed a system with pivots outboard of the fuselage (Polhamus and Hammond, 1967). The first production aircraft with swing wing capability was the F-111, which was developed in the 1960s and first entered service in 1967. The aircraft was designed to fulfill two roles: as a Navy fleet defense interceptor and an Air Force supersonic strike aircraft. The F-111 could take off and land within 2000 ft with fully extended wings and fly over Mach 2 with fully swept back wings. Its wing could sweep from 16 to 72.5 , although the inboard mounting of the wing pivots resulted in very high trim drag at supersonic speeds. Similar requirements also led TsAGI, the Soviet aerodynamics bureau, to explore the possibilities of variable geometry. TsAGI evolved two distinct planforms, differing mainly in the distance between the wing pivots; a wider spacing reduced the negative aerodynamic effects of changing wing sweep, and also provided a larger fixed-wing section that could be used for landing gear or pylons for stores. This could also be adapted to existing airframes, with the Sukhoi Su-17 (based on the earlier swept wing Sukhoi Su-7) and the Tupolev Tu-22M (based on the Tupolev Tu-22). However, the wide spacing also reduced the benefits of variable geometry. TsAGI also devised a vehicle with narrow pivot spacing, similar to that of the F-111. This design was used (albeit at different scales) for the MiG-23 fighter and the Sukhoi Su-24 interdictor, whose prototypes flew at the end of the 1960s, entering service in the early 1970s. The Northrop Grumman F-14 Tomcat entered service in 1974 and its wing sweep angle could vary between 20 and 68 to give the optimum lift-to-drag ratio in flight automatically. The F-14 can fly and land safely with asymmetrically swept wings in emergencies. Rockwell, meanwhile, adopted variable geometry for the Advanced Manned Strategic Bomber program that produced the B-1 Lancer bomber, intended to provide an optimum combination of high-speed cruising efficiency and fast, supersonic penetration speeds at extremely low level. Many military aircraft appeared with variable-sweep wings during the 1960s and 1970s, including the Panavia Tornado, the Mikoyan Mig-23, the Shukoi Su-22, the Su-24, and the Tupolev Tu-160 Blackjack, which first

flew in 1981. While variable sweep provides many advantages, particularly in takeoff distance, load-carrying ability, and in the fast, low-level penetration role, the configuration imposes a considerable penalty in weight and complexity. The advent of relaxed stability flight control systems in the 1970s negated many of the disadvantages of a fixed platform. The large gearbox that moved the wings of variable-sweep aircraft was complicated and heavy. Furthermore, all the aerodynamic loads on the wing are supported by the pivoting mechanism. The maintenance requirements therefore increased and fuel efficiency decreased. No new variable-sweep aircraft have been built since the Tu-160, though the replacement of the F-14 (the F/A-18E) has a reduced payload/range capability largely because of its small fixed wings. Changing wing sweep also moves the center of gravity, as well as the aerodynamic center, thus affecting the longitudinal stability of the airplane. The sweep angle also influences the lateral stability, and sweeping the wing back has a stabilizing effect similar to adding dihedral (Shortal and Maggin, 1946; Hunton and Dew, 1947; Kress, 1983). The oblique wing was first proposed by German designer Richard Vogt in 1942, for the Blohm & Voss P.202 variable-geometry jet fighter project. The pivoting oblique wing was mounted on top of the fuselage. NASA aerodynamicist Robert T Jones developed the theory for oblique-wing aircraft in 1952, and concluded that an oblique wing, supersonic transport aircraft might achieve twice the fuel economy of an aircraft sporting a conventional wing. By 1961, Handley Page designer Geoffrey Lee proposed a Mach 2 oblique flying-wing airliner, the Sycamore. Oblique wings have failed to make much progress, although several experimental aircraft have been tested. NASA Dryden flew a small, remotely piloted oblique-wing research aircraft in the mid-1970s and followed this with the single-seat AD1 with a fuselage-mounted oblique wing pivoted by up to 60 . A supersonic oblique-wing research aircraft based on NASAs Vought F-8 fly-by-wire test bed was proposed, but the program was canceled. NASA Ames completed a preliminary design study of a 500-seat, Mach 1.6 supersonic transport in 1991. The 124-m span aircraft was designed to take off and land with 37.5 oblique sweep, increasing to 68 in the cruise. NASA built a small-scale remotely piloted demonstrator that flew in 1994, and demonstrated stable, controlled flight over a range of sweep angles. Oblique-wing designs were ultimately rejected by NASAs High Speed Research program, but DARPA is now revisiting the concept. Challenges range from ground maneuvering, and the integration of fully embedded engines into the airframe, to the control of a tailless, unstable, variable-sweep flying wing at speeds up to Mach 2. The coupling between the asymmetric aircrafts aerodynamic

Morphing Aircraft

839

and aerostructural modes pose a control challenge. DARPA has awarded Northrop Grumman a contract for an X-plane Obliquing Flying Wing (OFW) demonstrator, known as the Switchblade, which is scheduled to fly in 2020. Sullivan and Watkins (2003) investigated the design of an obliquing wing aircraft (an aircraft capable of rotating its wing around a central pivot) also capable of both symmetric and asymmetric span extensions. The three proposed morphing goals were: symmetrical span extension, roll control using asymmetric span extension, and a combination of wing obliquing with asymmetric span deployment to trim the aircraft. The maximum rolling moment with asymmetric span actuation was found to be equivalent to the 5 aileron deflection. A prototype was designed adopting a worm gear system for the obliquing wing and two servos commanding a pulley and cable system to deploy the telescoping wing. No experimental prototype has been built. Marmier and Wereley (2003) built and tested in a wind tunnel a variable-sweep UAV. A pair of antagonistic inflatable bellow actuators were embedded in a cylindrical polyvinyl chloride fuselage and controlled by solenoid valves that allowed wing sweep to vary between 0 and 45 . Slide rods guaranteed a smooth translation. The wings had a NACA 0012 airfoil with 00 a 45 span and made with foam cores wrapped with fiber glass. Neal et al. (2004) provided a sweeping mechanism for their shape-morphing UAV that was actuated by means of two electromechanical, lead-screw actuators via a three-bar linkage. Neal et al. (2006) also implemented a variable-geometry tail and increased the structural strength. They included new direct-drive sweep actuators with a higher load capacity and added rotating sweep supports to transmit reaction torques from the twist actuation and normal wing loads. Grant et al. (2006) investigated miniature air vehicles with a variable-sweep wing inspired by birds for operation within urban environments at low altitudes. Agility, in terms of enhanced turning capabilities, and crosswind rejection are important features, and hence seagulls inspired the prototype. The inboard and outboard wing sections were capable of independent symmetric or asymmetric actuation. The wing sections were connected to the fuselage and each other through a system of spars and joints. The results demonstrate that symmetric sweep reduces the turn radius and the asymmetric sweep maintains sensor pointing in cross-winds. NextGen Aeronautics (Andersen et al., 2007; Flanagan et al., 2007) developed a UAV, called the MFX-1 and referred to as the BatWing, with a wing that could undergo significant sweep changes during flight. An electric motor deformed an endoskeleton wingbox structure that was covered with an elastomeric skin with out-of-plane stiffeners. Five distinct

configurations can be maintained in flight, and the aspect ratio can change by 200%, the span by 40%, and the wing area by 70%. Wing sweep and wing area could be controlled independently using four-bar linkages, revolute joints, and slider mechanisms. The skin could strain by over 100% but still withstand airloads of 400 lbs/ft. NextGen successfully flight tested the MFX-1 in 2006. NextGen then developed a larger morphing UAV named MFX-2 that flew in 2007 and demonstrated a 40% change in wing area, a 73% change in span, and a 177% change in aspect ratio. The scissor-mechanisms of Joo et al. (2006) or Bharti et al. (2007) also resulted in a change in sweep. The mechanism by Bharti et al. (2007) was based on the TSCh wing by NextGen Aeronautics Inc. (Andersen et al., 2007; Flanagan et al., 2007) and implemented as a pin-jointed mechanism. Grady (2010) extended the wing concept of NextGen Aeronautics, focusing on optimal actuator orientation for rigid and flexible structures. As the links flex, there is a non-linear stiffening effect and an increase in the bending of the flexible links further from the root of the wing. Yu et al. (2007) used SMP for the shear deformation of a scissor-mechanism. The twisting deformation of a rectangular composite plate made of SMP and carbon fiber upon heating could rotate a hinge in the mechanism. Mattioni et al. (2006) investigated a variable-sweep wing concept based on bi-stable composite spars. The wingbox in their design consisted of two spars with an interconnected trussrib structure. Each spar had a significant transverse curvature, which increased the bending stiffness and also allowed the spar to behave like an elastic hinge under high drag loads. The design could eliminate hinges and mechanisms, but could suffer from fatigue. No skin was included and adding a skin could interfere with the snapping motion. Sofla et al. (2010) proposed a wing that deforms to target shapes by actuating the wingbox using SMAs. The prototypes show excellent and smooth movement under representative loads. Aerodynamic analysis was conducted to evaluate the effect of reconfigured wing shape on the lift and drag coefficients. Variable-wing sweep has also been employed on MAVs, even though compressibility effects at low speeds and low Reynolds numbers are negligible. Hall et al. (2005) highlighted that the goal of variable sweep in this case is to increase the minimum drag speed. CFD analysis showed that the sweep angle decreased the liftcurve slope through a decrease in effective camber. The smaller frontal profile reduced the drag at high speeds. Summary All three contractors (Lockheed Martin, Hypercomp/ NextGen and Raytheon Missile Systems) of the recent (20002007) DARPA Morphing Aircraft Structures

840

S. BARBARINO ET AL.

(MAS) program conducted a functional analysis that concluded that changing wing planform area and wingspan were the primary enablers of a new class of morphing air vehicles. Although several parameters define wing planform (and can be sometimes interconnected), this review highlights that variable sweep has been implemented on real world fixed wing aircraft more than any other morphing strategy. Variable sweep was adopted for military fighter aircraft from the 1950s, mainly to achieve faster supersonic cruising speeds. In contrast, recent studies have focused on MAVs and UAVs, using variable sweep in conjunction with span to improve the aerodynamic efficiency and multi-role capabilities (to achieve both lowspeed loiter and high-speed dash). Most studies produced at least a bench prototype. Some studies validated their predictions with wind tunnel tests, for example the Agile Hunter manufactured by Northrop Grumman (Bye and McClure, 2007; Ivanco et al., 2007; Love et al., 2007). Only NextGen Aeronautics (Andersen et al., 2007; Flanagan et al., 2007) actually flew two UAVs (MFX-1 and MFX-2), demonstrating in-flight wing morphing capabilities which also affected sweep. All the solutions investigated so far adopt traditional actuation systems, with the exception of the SMP hinge proposed in the proof-of-concept prototype by Yu et al. (2007). Similarly, rigid or sliding skins are adopted in most solutions, with the exception of the flying test bed produced by NextGen, which had a flexible skin. Moreover, variable sweep is set to undergo further investigation in the near future under a different reincarnation, within the DARPA OFW program. Variable span is a wing planform-morphing capability that has been extensively investigated, with studies performed on both fixed and rotary wings, many of which have been developed to the stage of wind tunnel tests. Regarding rotary wings, the change in the blade span is mainly beneficial for tiltrotors, which have opposing requirements during vertical flight and cruise. For helicopters, span morphing also produces benefits in terms of flight envelope expansion. For fixed wings, span increase has been considered to improve cruise efficiency at low speeds without the typical drawbacks (structural and dynamical) of high-aspect ratio wings, or used as a means of roll control (when span increase is asymmetrical). Most of the solutions found in the literature adopt traditional actuators, mainly pneumatic, coupled to telescopic mechanisms, to reduce the weight penalty due to morphing. This is generally possible because most applications are for UAVs. The use of the available centrifugal force on rotor blades for actuation, as proposed by Prabhakar et al. (2007), is an interesting concept. The skin is generally considered to be rigid, composed of several sliding parts, although some authors also investigated the adoption of a flexible, stretchable rubber (with some limitations). Research on this topic is well

spread over recent decades, and there is no indication that this trend will stop in the near future. Finally, variable chord has been investigated least, with only a few works that are still at the early stages of development, and mainly based on small-scale aircrafts (UAVs) or rotor blades. The limited investigation arises because the small benefits and large challenges involved in such a solution. Increasing the chord of a wing (particularly for a fixed wing) represents a big challenge from a structural perspective (for instance, how to move the spars or how to stretch the skin) and still produces small advantages when compared to other morphing configurations. However, the possibility to increase wing area or to combine a chord increase with variable camber of the airfoil could lead to morphing solutions capable of further improving performance and competing with high-lift devices currently adopted on widebody aircraft: the latter solution may only require a limited part of the wing chord to undergo morphing, thus making such a configuration easier to implement from a structural perspective. WING OUT-OF-PLANE TRANSFORMATION Wing out-of-plane transformation is mainly affected by three parameters (individually or in combination): twist, dihedral/gull, and span-wise bending. Rigid-body rotation of the aerodynamic surfaces is also considered in this section if the rotation is induced by smart materials. Tables 48 show the papers for each geometrical parameter: due to the large literature on twist, an additional categorization has been adopted based on the activation method/strategy. Twist Using Active Aeroelastic Concepts Varying the twist distribution of the wing to enhance flight performance and control authority of the air vehicle can be regarded as the oldest form of morphing. The Wright Brothers employed the wing warping technique to change the twist of a flexible wing and provide roll control for their first flying machine (Pendleton, 2000; Pendleton et al., 2000). However, the quest for enhanced performance and higher airspeed prohibited designers required stiff structures to avoid aeroelastic instabilities and to meet loading requirements. This increased the structural weight of the vehicle and penalized its performance. Advances in aerospace materials and the continuous pressure to enhance aircraft performance and expand their flight envelope have focused the interests of aircraft designers again on using the flexibility of the structure in a beneficial manner. Rockwell International pioneered the Active Flexible Wing (AFW) program in the 1980s (Miller, 1988), which exploited wing flexibility to reduce structural weight and prevent roll control problems (aileron reversal) for

Morphing Aircraft Table 4. Out-of-plane morphing: Twist.


Vehicle Geometrical parameters Twist Twist Twist Twist Twist Investigation

841

Experimental References Miller (1988) Griffin and Hopkins (1997) Chen et al. (2000) Nam et al. (2000) Pendleton et al. (2000) Wilson (2002) Clarke et al. (2005) Kuzmina et al. (2002) Amprikidis et al. (2004) Cooper et al. (2005) Amprikidis and Cooper (2003) Cooper (2006) Florance et al. (2004) Runge et al. (2010) Garcia et al. (2003) Stanford et al. (2007) Bartley-Cho et al. (2004) Guiler and Huebsch (2005) Majji et al. (2007) Mistry et al. (2008) Vos et al. (2008b, 2010) Mistry et al. (2010) Category Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Size Advanced fighter Fighter Advanced Fighter Fighter Advanced fighter Actuation Hydraulic Skin Rigid Rigid Electric motor Mechanical Hydraulic Rigid Flexible Rigid Purpose Control (Roll) Control (Roll) Control (Roll) Control (Roll) Control (Roll) Numerical  8    Prot.  8 8 8  WT  8 8 8  FTB 8 8 8 8 

Twist

Fixed wing

Pneumatic

Rigid

Control (Yaw)

Twist

Fixed wing Fixed wing Advanced Fighter

Twist Twist Twist

Electric MotorPneumatic Hydraulic

Rigid

Performance (CD) Control (Roll)

  

   

  8 

Rigid

8 8 

Fixed wing

MAV

Servo

Flexible

Control (Roll)

Twist Twist Twist Twist Twist Twist

Fixed wing Fixed wing Fixed wing Rotary wing Fixed wing Fixed wing

UCAV UAV

Ultrasonic Motor Servo Servo

Stretchable Flexible Elastomeric

Control (Roll and Pitch) Control (Roll, Yaw, Pitch)

8 8 

   8  

   8  8

8 8 8 8 8 8

Helicopter Threaded rod Threaded rod Flexible Rigid

Performance Control (Roll) Performance

  

Helicopter

high-performance fighter aircraft at large dynamic pressures. In the AFW technology, the conventional control surfaces were not used as primary control force-producing devices, but were used as aerodynamic tabs to control the aeroelastic twist of the wing. The wing aeroelastic twist was used to produce the required roll moments for control. This allowed the aircraft to operate beyond the dynamic pressure where conventional aileron reversal begins. The AFW program was followed by the Active Aeroelastic Wing (AAW) research program, funded jointly by the US Air Force and NASA. In the AAW program, the AFW technology was

implemented and tested using a modified F/A-18 fighter to demonstrate its feasibility (Pendleton, 2000; Pendleton et al., 2000; Wilson, 2002). The F/A-18 wing skin panels near the rear section of the wing just ahead of the trailing edge were replaced with thinner, more flexible skin panels. The leading-edge flap was split into separate inboard and outboard segments, and additional actuators were added to operate the outboard leading-edge flaps separately from the inboard leading-edge surfaces. Trailing-edge control surfaces and leading-edge flaps were used to control the aeroelastic twist of the wing and to achieve the required rolling power. Flight tests

842

S. BARBARINO ET AL.

Table 5. Out-of-plane morphing: Twist and rigid-body rotation with piezoelectric actuation.
Vehicle Geometrical parameters Bending and twist Twist Camber and twist Rotation Investigation Experimental References Crawley et al. (1989) Barrett (1990, 1992a, b) Ehlers and Weisshaar (1990, 1992) Barrett et al. (1997) (Mothra) Barrett et al. (1996) (Gamara) Barrett et al. (1998) (Kolibri) Barrett and Stutts (1998) Barrett et al. (1998, 2001) (LuMAV) Sahoo and Cesnik (2002) Cesnik and Brown (2003) Detrick and Washington (2007) Category Fixed wing Rotary wing and missiles Fixed wing Size Actuation PZT PZT PZT Skin Flexible Purpose Performance Numerical    Flexible Control (Yaw and pitch only) Control         Prot.   8 WT   8 FTB 8 8 8

Fixed wing

Small UAV

PZT

     8 8 8

     8 8 8

   8  8 8 8

Twist

Rotary wing

Small UAV

PZT

Rigid

Rotation

Rotary wing (tethered) Missiles Rotary wing

MAV

PZT

Rigid

Control

Rotation Rotation

PZT MAV PZT

Rigid Rigid

Performance (CL) Control

Twist Twist Twist

Fixed wing Fixed wing Fixed wing

UCAV Joint-wing Sensorcraft UAV

PZT PZT PZT

Flexible Flexible FlexibleSliding

Control (roll) Control (roll) Control (roll)

demonstrated that 1g/360 rolls, and rolling pullout maneuvers were possible within the structural and hinge-moment limits of the aircraft (Clarke et al., 2005). Active aeroelastic structures (AASs) are receiving much interest for aeronautical applications due to the potential to improve drag performance, as well as roll and loads control. They are a sub-set of adaptive structures that allow significant performance and control improvements by manipulating the aerodynamic shape of a lifting surface by modifying the internal structure, without the need for large planform modifications that typically use complex and heavy mechanisms. A variety of AAS concepts has been developed and studied to control the aeroelastic twist of lifting surfaces. For example, Griffin And Hopkins (1997) investigated the use of the smart spar concept to vary the torsional stiffness and control the aeroelastic behavior of the wing to enhance the roll performance of high-performance aircraft at large dynamic pressures. The solution proposed was based on the simultaneous actuation of control surfaces and the modification of the wing torsional stiffness using the smart spar concept. The smart spar concept has a web that can either transmit shear between its

upper and lower caps or disable the shear transmission between the upper and lower caps. This is achieved by adjusting the position of the smart spar between its original position (along the leading edge) and a new position where it runs from the leading-edge root to the trailingedge tip of the wingbox. Chen et al. (2000) developed the variable stiffness spar (VSS) concept to vary the torsional stiffness of the wing and enhance the aircraft roll performance. Their VSS concept consisted of a segmented spar having articulated joints at the connections with wing ribs and an electrical actuator capable of rotating the spar through 90 . At the horizontal position, the segments of the spar are uncoupled and the spar offers no bending stiffness. While at the vertical position, the segments join completely and the spar provides the maximum torsional and bending stiffness. The concept allows the stiffness and aeroelastic behavior of the wing to be controlled as a function of the flight conditions. Nam et al. (2000) advanced the VSS concept and developed the torsionfree wing concept, where an enhanced VSS wing design is used to attain further post-reversal aeroelastic amplification. The primary structure of the torsion-free wing

Morphing Aircraft Table 6. Out-of-plane morphing: Twist and rigid-body rotation with SMA actuation.
Vehicle Geometrical parameters Twist Twist Twist Twist Twist Twist Twist Twist Twist Investigation

843

Experimental References Martin et al. (1998) Chandra (2001) Prahlad and Chopra (2001) Nam et al. (2002) Prahlad and Chopra (2002) Elzey et al. (2003) Ruggeri et al. (2008a, b) Sofla et al. (2008) Pagano et al. (2009) Category Fixed wing Rotary wing Rotary wing Fixed wing Rotary wing Fixed wing Rotary wing Fixed wing Rotary wing Size Advanced fighter Helicopter Tiltrotor Fighter Tiltrotor Actuation SMA SMA SMA SMA SMA SMA SMA SMA SMA Sliding Flexible Flexible Control (roll) Performance Skin Rigid Purpose Performance ( C L) Numerical       8   Prot.    8      WT  8 8 8 8 8  8 FTB 8 8 8 8 8 8 8 8 8

Tiltrotor

Performance

Helicoptor

(References: Caldwell et al. (2007); Geometrical parameter: Twist; Category: Rotary wing; Actuation: SMA; Purpose: Performance; Investigation: Numerical)

Table 7. Out-of-plane morphing: Dihedral/gull.


Vehicle Geom. parameters Category Gull Fixed wing Flying wing Fixed wing Fixed wing Fixed wing Fixed wing UAV Investigation Experimental References Abdulrahim and Lind (2004) Size MAV Actuation Linear actuator Servo Skin Flexible (nylon film) Purpose Numerical Prot.         8 WT   8 FTB  8

Bourdin et al. Dihedral (2008, 2010) Gatto et al. (2010) Shelton et al. Dihedral (2006) Bye and McClure (2007) Supekar (2007) Ursache et al. (2007, 2008) Dihedral (wing folding) Dihedral Dihedral

UCAV

Electrical actuator Servo Servo

Control (Flight dynamics) Flexible Control (carbon fiber) (multi-axis control) Performance (range, endurance, L/D) Elastomer Performance

  

 8 8

 8 8

UAV Narrow body transport aircraft

Performance Corrugated Performance (SAR)

Cuji and Garcia (2008) Smith et al. (2010)

Dihedral Dihedral

Fixed wing Fixed wing

Narrow body transport aircraft

Control (roll  and turning flight) Performance  (SAR, take-off, and landing distances)

8 8

8 8

8 8

consists of two parts. The first part is a narrow wingbox tightly attached to the upper and lower wing skins to provide the needed basic wing torsional stiffness. The second part consists of two VSSs placed near the leading and trailing edges, passing through holes in all of the rib. This torsion-free wing can provide significant aeroelastic amplification to increase the roll-rate between 8.44% to 48% above the baseline performance. Florance et al.

(2004) extended the use of the VSS concept to exploit wing flexibility and improve the aerodynamic performance of the vehicle. Their wing incorporated a spar with a rectangular cross-section that runs from the wing root up to 58% of the wingspan. In Europe, the Active Aeroelastic Aircraft Structures (3AS) research project, which involved a consortium of 15 European partners in the aerospace industry and was

844

S. BARBARINO ET AL.

Table 8. Out-of-plane morphing: Span-wise bending.


Vehicle Geometrical parameters Span-wise bending Span-wise bending Investigation Experimental Authors Wiggins et al. (2004) Manzo et al. (2005) Manzo and Garcia (2010) Lazos and Visser (2006) Sofla et al. (2010) Category Fixed wing Fixed wing UAV SMA and DC motor Stretchable Size Actuation Skin Purpose Performance Performance (Drag) Numerical   Prot.   WT 8  FTB 8 8

Span-wise bending Span-wise bending

Fixed wing Fixed wing UAV SMA Flexible

Performance (L/D) Performance (L/D)

 

 

 8

8 8

partially funded by the European Community, developed active aeroelastic design concepts to exploit structural flexibility in a beneficial manner to improve the aircraft efficiency (Kuzmina et al., 2002; Simpson et al., 2005). One of these novel concepts is the allmoving vertical tail (AMVT) with a variable torsional stiffness attachment. The AMVT concept provides a smaller and lighter fin while maintaining stability and control effectiveness at a wide range of airspeeds. The AMVT employed a single attachment and the position of the attachment can be adjusted in the chordwise direction relative to the position of the center of pressure to achieve aeroelastic effectiveness above unity (Amprikidis et al., 2004). The 3AS project also investigated a variety of variable stiffness attachments and mechanisms for the AMVT concept, including a pneumatic device (Cooper et al., 2005). Amprikidis and Cooper (2003) and Cooper (2006) investigated two AAS concepts that modified the static aeroelastic twist of the wing by modifying its internal structure. The first concept exploited the chord-wise translation of an intermediate spar in a three-spar wingbox to vary its torsional stiffness and the position of the shear center. The second concept was similar to the VSS concept and used rotating spars to vary the torsional and bending stiffnesses, and also the shear center position. Prototypes of the concepts were built and tested in a wind tunnel to examine their behavior under aerodynamic loading. Ajaj et al. (2011a, b) investigated the adaptive torsion structure (ATS) concept to control the static aeroelastic twists of the wing. Their concept has a wingbox with two spars, where the webs of the spars can translate inward to vary the torsional stiffness and the position of the shear center. This allows the external aerodynamic flow to induce aeroelastic twist deformation on the

wing that can be used in a beneficial manner to enhance the performance or control authority of the air vehicle. Garcia et al. (2003) and Stanford et al. (2007) investigated the roll control of a MAV by twisting its flexible wing. In their design, torque rods actuated by a single servomotor created equal and opposite deformations for each wing. Their numerical analysis showed that the wing morphing could roll the MAV but led to a considerable drag penalty. The DARPA Smart Wing program phase II (BartleyCho et al., 2004) demonstrated the gradual changing of the airfoil camber to create wing twisting. In their design, the control surface was sectioned into 10 segments. Each segment could undergo out-of-plane shape changes by means of an eccentuator. The eccentuator was a concept developed by Vought (Musgrove, 1976, 1981) in the late 1970s for variable-camber control surfaces for use on the supersonic transport. It consists of a bent beam that converts a rotary input motion into a vertical and lateral translation. The vertical motion can then be delivered to the structure to flex it. Each segment in the final design was connected to an eccentuator driven by an ultrasonic motor. A continuous silicone skin on the top and bottom surfaces covered the segments. Guiler and Huebsch (2005) developed an effective wing morphing mechanism to vary wing twist to control a swept wing tail-less aircraft. A torque rod mechanism was located at the leading edge and 10 freefloating wing sections between the tip rib wing section and the wing root were connected at the trailing edge by a spring-loaded cable. The assembly was covered with a latex skin. Initial wind tunnel testing indicated that this morphing mechanism provided adequate control forces and moments to control a UAV and potentially a manned aircraft. This morphing wing

Morphing Aircraft

845

mechanism could allow a swept wing tail-less aircraft to fly in cruise without the added drag of built in washout or winglets normally used to prevent adverse yaw during maneuvering. Tests showed that lift-to-drag ratio could be improved by 15% compared to an elevon-equipped wing with built-in 10 of washout. Majji et al. (2007) studied a wing consisting of an elastic wingbox structure (ABS plastic material) covered with an elastomeric skin. The wingbox was rigidly coupled to four concentric tubes, which were independently attached to the wing at four locations along the span. The outer tubes passed through the inner tubes and were connected to servomotors at the wing root. The wing could be twisted by the arbitrary rotation of the tubes. It was shown that the angle of attack envelope of the twisted wing was increased because the tip section stalled later than the inner sections. Vos et al. (2008b, 2010) investigated a novel mechanism to actively control wing twist. A longitudinal slit was introduced along the entire span of the wing, and a threaded rod mechanism was positioned near the slit to induce relative displacement between the upper and lower skin. This induced warping of the wing and hence modified its twist distribution. Mistry et al. (2008) examined the twisting of an Ibeam spar through the application of differential loads (Vlasov Bimoments) applied at the tip of a variable-twist helicopter rotor blade. By clamping the web of the spar at the root, while leaving the flanges free to warp, the Ibeam will have no warping restraint, and displays the reduced torsion stiffness of a freefree beam, producing larger twist deformations. Mistry et al. (2010) examined a warp-induced twist concept to achieve a large quasistatic twist of a helicopter or tilt-rotor blade. The concept employed a cylindrical spar with rotating ribs with the ribs attached to the skin, which was slit along the trailing edge. Warping the skin then produced twisting of the blade section. A threaded rod near the trailing edge implemented warp actuation. During warp actuation, the torsion stiffness reduced and the blade twisted easily. However, without power, the threaded rod effectively produced a closed section with high torsional stiffness. A prototype based on this concept was built and tested, and the results show tip twist variations of up to 18 . Runge et al. (2010) developed a novel method to control the wing twist by modifying the internal structure of the wing using a mechanical system to change the shear center position relative to the aerodynamic center, although details of the internal mechanism were not provided. Their method is similar to the approach used by Amprikidis and Cooper (2003) and Cooper (2006). A full demonstrator was built and tested to verify the ability of the concept to control the twist of the structure.

Piezoelectric Actuation of Twist and Rigid-Body Rotation Piezoelectric actuators are widely used in a variety of smart structure designs for aeronautical applications due to their high bandwidth, high output force, compact size, and high power density properties. Crawley et al. (1989) presented one of the first investigations of active aerodynamic surfaces. They constructed a bending-twist coupled graphiteepoxy plate actuated by conventionally attached piezoceramic sheets. When the plate was subjected to airloads, it bent further, which increased the twist and hence the control deflections were effectively magnified through the coupling. The first twist-active piezoceramic-actuated missile wing and helicopter rotor blades were designed and prototyped from 1989 (Barrett, 1990). Ehlers and Weisshaar (1990, 1992) presented research on adaptive wings and missile fins and showed that aeroelastic tailoring coupled with directionally attached piezoceramic (DAP) elements could be used to generate sufficient deflections for flight control. The first torque-plate rotor generated 4.5 static pitch deflections and was capable of 3/rev individual blade control (Barrett, 1992a, b). The torque-plate rotor concept was matured in the following years leading to a whirling stand test of a 120-cm solid-state adaptive rotor. This test showed that 8 static pitch deflections could be generated and dynamic pitch deflections could be commanded up to 2.5/rev (Barrett and Stutts, 1997). Barrett and Law (1998) designed, fabricated, and tested a new class of twist-active wing where the pitch angle of the wing root was adjusted using a DAP torque-plate. Wind tunnel tests showed that the wing tip of their prototype could twist up to 1.5 which is comparable to 8.4 of aileron deflection. The root-twist concept was also over 20% lighter than conventional ailerons. During the same period, the Flexspar solid-state adaptive stabilator was introduced (Barrett, 1996). The Flexspar actuator was used in several vehicles including the first full missile configuration with an adaptive wing. Even without aeroelastic tailoring, deflections in excess of 30 were achieved in some configurations. The Flexspar configuration employs a high-strength main spar around which a flexible skin structure (shell) is pivoted. A piezoelectric bimorph actuator is mounted within the flexible skin structure and causes the skin structure to rotate about the spar. Due to aerodynamic loading, the deformation of a Flexspar-actuated control surface is a combination of rigid-body rotation, twist and some change in camber. The first fixed-wing UAV to fly using adaptive materials for all flight control, Mothra, was flown in 1994. This vehicle employed Flexspar actuators for control of rudder and elevator and it had no other control surfaces. Although the two Flexspar-actuated stabilizers were not completely solid-state, the aircraft showed the feasibility of piezoelectrics for flight control (Barrett et al., 1997). The first

846

S. BARBARINO ET AL.

rotary-wing UAV to fly using adaptive materials for all flight control, Gamara, was flown in 1996 (Barrett et al., 1998). A commercially available radio-controlled rotarywing UAV was stripped of all conventional servoactuator related hardware, including the servopaddle assembly, and replaced with a pair of two solid-state adaptive rotor torque-plate servopaddles. Flight tests showed the feasibility of the piezoelectric control and benefits in drag (26% reduction), weight (8% total gross weight reduction), and part count (reduced from 94 to 5). In 1997, Kolibri became one of the first rotarywing MAVs to use piezoelectric actuators (Flexspar) for flight control. This tethered, 15-cm diameter vertical takeoff and landing (VTOL) aircraft demonstrated controlled flight with the Flexspar actuators (weighing 5.2 g each) and achieved 11 static stabilator pitch deflection (Lee, 1997; Barrett and Howard, 2000). In 1999, the Lutronix Coleopter VTOL MAV (LuMAV) followed Kolibri by replacing the electric motor with an internal combustion engine and removing the electrical tether (Barrett and Lee, 2000; Barrett et al., 2001). This rotary-wing aircraft also used the Flexspar actuators. Barrett and Stutts (1998) proposed an actuator that uses pairs of piezoceramic sheets arranged in a pushpull assembly to turn a spindle of an aerodynamic control surface. To demonstrate the actuator concept, a maximum compression design for an air-to-ground weapon was conceived. Modeling of the actuators was accomplished by laminated plate theory and kinematics. A 50% scale experimental model was built for bench and wind tunnel testing. Bench testing showed that actuators could generate 10 fin deflections with a corner frequency of 59 Hz with good correlation between theory and experiment. Wind tunnel testing showed the feasibility of the concept. Sahoo and Cesnik (2002) investigated the use of high authority anisotropic piezoelectric actuators to induce wing warping and achieve roll controllability instead of traditional discrete control surfaces for the next generation of unmanned combat air vehicle (UCAV). A numerical design environment was developed for the integrally distributed anisotropic piezocomposite actuators in a composite wing. A Boeing X-45A-based UCAV model was used to highlight that the actuators not have sufficient control authority to achieve a representative roll rate while satisfying other design constraints. A three- to four-fold increase in actuator authority is needed for this type of aircraft and the improved properties of single-crystal fiber composites may provide the desirable authority. Similar results were obtained by Cesnik and Brown (2003) for the use of high anisotropic piezoelectric actuators (APA) to induce wing warping actuation for roll control of the joined-wing Sensorcraft. Detrick and Washington (2007) developed two designs for morphing wings for UAVs. The first design split the supporting

ribs and achieved wing twist using levers and a piezoelectric actuator. This design suffered from limitations because the wing skin was attached directly to the ribs and so stiffened the wing. The second concept used onepiece ribs interconnected by a truss structure. Wing twisting was accomplished by varying the length of some of the truss structure members. The ribs had rollers in contact with the skin to allow sliding, and so avoid stiffening the wing. Wilbur and Wilkie (2004) proposed an active twist rotor (ATR) concept with active fiber composites piezocomposite actuators. The concept can reduce vibration and noise, and reduce the rotor power requirements by 23% and improve its performance by 1%. Shape Memory Alloy Actuation of Twist and Rigid-Body Rotation The Smart Wing program of DARPA, AFRL, NASA, and Northrop Grumman (Martin et al., 1998; Kudva and Carpenter, 2000) realized an UCAV by integrating SMAs within a morphing wing. This experimental study implemented both shape memory alloy (SMA) torque tubes for wing twist and flexible leading and trailing edges actuated by SMA wires. The SMA torque rods were connected to a torque transmission rod to transfer the loads to the wingbox. Martin et al. (1998) provided a detailed description of the design, integration, and testing of the mechanism. Barrett et al. (2001) introduced the pitch active SMA wing UAV with a 2 m span. This flight-tested fixed-wing aircraft had SMA wires that changed the pitch of the main wings. The SMAs caused a rigid-body rotation of the wings. Nam et al. (2002) used SMA spars to replace the mechanically actuated VSS concept to enhance the aeroelastic performance of the wing. They adopted active property tuning where the SMA was embedded without plastic elongation. Therefore, no internal stresses are induced, and the SMA varies its stiffness. A wing model based on the F-16 demonstrated that the SMA spar can further amplify the aeroelastic forces and significantly enhance roll performance compared to the traditional VSS concept. Furthermore, locating the SMA spar toward the trailing edge maximized the improvement in roll rate in comparison to the traditional VSS concept (61%). Elzey et al. (2003) developed an antagonistic shape morphing structural actuator, comprising a cellular flexible core sandwiched between SMA face sheets that induced curvature upon heating. The core was an assembly of modular elements able to rotate relative to one another (vertebrate structure); hence, the actuator could reverse its shape without any bias mechanism to provide the elastic restoring force. A prototype wing was built in which the vertebrate actuators were incorporated as ribs. The wing also twisted by asymmetric actuation of

Morphing Aircraft

847

its two SMA-actuated vertebrate beams. Similarly, Sofla et al. (2008) used antagonistic flexural cells to create two-way SMA flexural actuators using a one-way shape memory effect. The cells provided four distinct positions for each segment of the actuator, and two of the positions required no external energy to be maintained. These actuators could control the twist of wing sections if they are incorporated as ribs and are actuated asymmetrically. Lv et al. (2009) developed an SMA torsion actuator based on NiTi wires and a thin-walled tube for an adaptive wing demonstration system. The angle of attack of the wing could be changed continuously up to 15 in 1 s. A novel smart wing rib with the swing angles up to 10 was investigated. SMAs have also been applied for wing twist on helicopter rotor blades. Prahlad and Chopra (2001) developed and tested a torsional SMA actuator to alter the twist distribution for a tiltrotor blade between hover and forward flight. The design of the actuator (torque tube) was optimized as a function of heat transfer, temperature, and actuation recovery torque. The effect of heat treatment on the tuning of the actuation characteristics of the SMA tube was also investigated, and thermoelectric modules were used to release excess heat through the blade surface to improve the cooling and reduce the actuation time. Chandra (2001) developed a method to induce twisting deformation in a solid-section composite beam using SMA bender elements. Bendingtorsion coupled graphiteepoxy composite beams with Teflon inserts were built using an autoclave molding technique and the Teflon inserts were replaced by SMA bender elements. The composite beams were actuated by heating the SMA via electrical resistive heating, and the bending and twisting deformation measured using a laser system. Such beams with SMA elements can be used as spars in rotor blades to induce twist. The goal of the reconfigurable rotor blade program funded by NAVAIR was to demonstrate the potential to improve rotorcraft performance by optimizing the configuration of major structures in flight (Bushnell et al., 2008; Ruggeri et al., 2008b). The SMA actuation system employed 55-Nitinol (Ni-55Ti) rotary actuators integrated as structural elements in a quarter-scale rotor blade to control twist. A passive torque tube transmitted the torque from the actuator assembly near the blade root to the tip of blade causing the blade to twist. A lightweight spring mechanism, the Strain Energy Shuttle, provided an energy storage element between the SMA actuator and the passive torque tube, and halved the actuator system weight (Calkins et al., 2008). The actuator was tested in 2007 using a scale three-blade hub assembly mounted on the Boeing Advanced Rotor Test Stand. The wind tunnel test was a high-fidelity assessment of the SMA actuator (Ruggeri et al., 2008a) and represents one of the first attempts to

produce a high-torque SMA actuator for the rotor environment. The actuator provided approximately 250 twist transitions during 75 h of testing with no loss of performance or operational anomalies. Pagano et al. (2009), within the European Friendcopter project, investigated the control of a rotor blade twist using an SMA-based device to extend the flight envelope of the helicopter. An SMA rod was pre-twisted to achieve a martensitic phase and then integrated in the span-wise direction within the blade structure at different positions. When heated, the SMA actuator transmitted a torque couple which induced twist into the blade. A prototype was been built and tested, and achieved a maximum angular rotation of 6 , with a corresponding transmitted moment of 21.6 Nm. Dihedral/Gull Recently variable dihedral/gull wings have aroused much interest because of their ability to enhance aircraft performance and flight control. A variable dihedral wing can: control the aerodynamic span; replace conventional control surfaces; enhance the agility and flight characteristics of high-performance aircraft; reduce the induced drag (by changing the vorticity distribution); and improve the stall characteristics. The IS-1 fighter, designed by Nikitin-Shevchenko in 1932, was one of the first applications of variable dihedral wings, and was capable of out-of-plane morphing from a bi-plane to a monoplane to operate at high speed. The XB-70 supersonic bomber also used a form of threedimensional wing morphing. This design used outer wing panel rotation to control the lift-to-drag ratio at both low subsonic and supersonic speeds. Abdulrahim and Lind (2004) studied the flight dynamics and characteristics of a variable gull-wing morphing MAV. They used a hinged spar structure to split the wing into inboard and outboard partitions and vary the dihedral angle of each partition. A vertical linear actuator controlled the angle of the inboard spar via a telescoping shaft to prevent binding. The angle of the outboard spar was passively controlled via a mechanical linkage connected to the fuselage. During actuation, the linkage caused the inboard and outboard sections to deflect in opposite directions. Shelton et al. (2006) studied the benefits of active multiple winglets for a UAV. The use of actively controlled winglets can enhance the low-speed performance and maneuverability of the vehicle and can increase the range and endurance of the vehicle by up to 40%. As part of the MAS program, Lockheed Martin developed the folding wing (Z-wing), where the span length, aspect ratio, and effective sweep angle may be varied (Skillen and Crossley, 2008a). The folding wing design incorporates hinged joints at two span-wise

848

S. BARBARINO ET AL.

stations enabling rigid body motion of four primary wing sections. Two approaches to fold the wing were investigated: a thermo-polymer actuator driving a helical spline gear or electro-mechanical rotary actuators. However, the helical spline approach was high risk and hence electrical actuators were used (Bye and McClure, 2007). Elastomeric skins covered the entire wing and provided smooth shape changes. A morphing UAV aircraft was successfully flight tested. Bourdin et al. (2008) and Gatto et al. (2010) investigated the use of variable-cant angle winglets for morphing aircraft control. A pair of winglets with adaptive cant angle were mounted at the tips of a flying wing and actuated independently. A prototype of the concept was constructed with servo-driven winglets at the tips of a trapezoidal planar wing. A single pair of adaptive winglets cannot replace all the conventional control surfaces and an elevator was included to trim the aircraft, especially during level turns at arbitrary bank angles. The concept was modified by adding a second pair of folding wingtips (Bourdin et al., 2007, 2010), where the split wingtips are actuated independently. The concept is more effective at moderate and high lift coefficients, and hence could be applied to low-speed morphing aircrafts. Cuji and Garcia (2008) studied the dynamics of aircraft turning for symmetric and asymmetric V-shaped changing wings. Wings with asymmetric dihedral perform better in terms of bank angle, load factor, and rolling moment coefficient than wings with symmetric dihedral. Moreover, asymmetric dihedral can enhance the turning performance significantly without losing the ability to roll the vehicle. Very large dihedral angles are also good for flight missions where the bank angle is more important than the turning rate and radius. Ursache et al. (2007) investigated the MORPHing wingLET (MORPHLET) concept to enhance the flight performance of narrow body aircraft. The winglet system consisted of four partitions at the outboard region of the wing. The cant, twist, and span of each partition were adjusted to maximize the vehicle performance across the entire flight envelope. Low-fidelity mulitdisciplinary design optimization (MDO) studies were performed using the winglet system to maximize the specific air range (SAR) at three different points of the flight envelope: start of initial cruise, start of final cruise, and end of descent. Preliminary results indicated that a significant increase in SAR can be achieved. A mechanical prototype of the MORPHLET concept was designed, built, and tested (Ursache et al., 2008). A corrugated skin was incorporated to allow the relative displacement of the MORPHLET partitions. Smith et al. (2010) formulated a constrained multi-objective MDO analysis of the MORPHLET concept to quantify operational performance benefits, and showed that a 45% improvement in SAR over that for fixed winglets may be achieved.

Span-wise Bending Another way to achieve a continuous out-of-plane wing morphing is through the biologically inspired hyper elliptic cambered span (HECS) concept, developed by NASA researchers. Manzo et al. (2005) investigated two approaches to allow shape morphing of a furled HECS wing. The first approach employed a tendon-based spool system that was actuated using a DC motor, while the second approach used an SMA system. Manzo and Garcia (2010) studied the aerodynamic benefits of a furled and a planar (rigid) HECS wings, in contrast to an elliptical rigid wing. A wing was constructed that could mimic a furled HECS profile by splitting the wing into five segments across the span. The shape morphing of the HECS wing was achieved using SMA-actuated segments that were powered in tandem using closed-loop feedback control. Wiggins et al. (2004) investigated the feasibility of a singledegree-of-freedom mechanism to continuously morph a flat wing to a non-planar shape. Their scissor-like mechanism used a repeating quaternarybinary link configuration to translate the motion of one wing segment to the next. The mechanism was synthesized such that only one actuator displacement to the first linkage was required to deform all the segments. Sofla et al. (2010) developed a morphing wing concept where the wing can flex laterally in a continuous fashion and is similar to the HECS wing concept. A prototype, made from carbonepoxy and actuated using SMAs, was built to examine the performance of the actuators. The prototype showed excellent and smooth movement under applied representative loads. A three-dimensional, parametric aerodynamic analysis was also conducted to evaluate the effect of the reconfigured wing shape on the lift and drag coefficients. The power consumption of a typical UAV was estimated from the aerodynamic performance of the wing. The numerical results showed considerable benefits of the wing morphing concept. Supekar (2007) designed a morphing wing concept with adjustable span and gull angles. The gull angle was controlled using a servomotor and bell-crank arrangement, whereas the outer wing allowed an increase in span. Summary Out-of-plane morphing is probably the least common type of morphing solution, perhaps with the exception of wing twist. Twist is the oldest shape morphing, but was discarded for almost 80 years to prevent aeroelastic problems. Advances in aerospace materials (composites) have made twist morphing possible, for example with the AFW research program. Following the AFW program, large numbers of research projects across the world have investigated methods and techniques to

Morphing Aircraft

849

achieve twist morphing. The main reason for this large interest is that twist morphing can produce a significant impact on the aerodynamic behavior of a lifting surface without the need for large platform modifications, such those associated with variable sweep or span that usually require complex and heavy mechanisms. In addition, twist morphing (similar to camber morphing) can serve multiple tasks simultaneously, such as alleviate gust and maneuver load; increase the lift coefficient; and replace conventional control surfaces. Furthermore, various actuation methods ranging from SMA, piezoelectric, AAS, and others have been investigated for twist morphing. However the state-of-art piezoelectric actuators still fail to deliver the actuation authority required to provide adequate roll control for fixed-wing air vehicles. Most of the research has been focused so far on actively tailoring the wing elasticity to achieve desired twist; moreover, traditional actuation systems have been adopted in most of the studies which reached the wind tunnel test stage. Fewer studies have focused on dihedral/gull or spanwise bending morphing, although some have reached an advanced stage, even producing a successful flying test bed. The interest in these morphing capabilities mainly arose from the need to optimize the performance of fixed winglets. A variety of studies and research projects focused on variable-cant winglets to maximize the performance of the air vehicle or to replace conventional control surfaces. Actuators considered in these works are generally traditional. The most recent studies have focused on continuous geometric morphing of the wing (root to tip) to generate improved outcomes; however, this requires complex and heavy mechanisms coupled with flexible skins.

by conventional or smart material systems) can be distributed or localized. In this section, the main focus is applications of camber control in fixed-wing and rotary aircraft in the subsonic regime. The section is separated into four main parts. The first part is organized from a flight regime standpoint; therefore, examples of patents, flapping wing aircraft, wind mills, etc. (in addition to fixed wing and rotary aircraft) are given to better explain the motivation and the necessity of camber morphing. The second part deals specifically with fixed-wing and rotary aircraft in the subsonic regime and actuated with conventional systems. The third part deals with fixedwing and rotary aircraft (also in the subsonic regime) that are actuated with SMAs or similar muscle-like devices. In the final, fourth part, piezoelectric actuators are considered. Camber Motivation for Adoption on Fixed Wings Although the idea of changing the wing camber was born with the first airplanes, it is far from simple-todesign devices capable of achieving the necessary deformation and suitable control systems. Many patents have been obtained and research carried out in this regard. The need for wing curvature change arises from the possibility, in the subsonic regime, to adjust continuously the airfoil geometry at different flight conditions, thus increasing the lift/drag ratio (Spillman, 1992). The more possible intermediate configurations, the greater are the aerodynamic benefits. Some applications also realized a small curvature change on the wing upper surface to increase the aircraft performance in the transonic regime (Stanewsky, 2001). Parker (1920) patented one of the first examples of a variable-camber wing. This concept involved changing the wing configuration through aerodynamic loads on the wing. This scheme divided the wing into three sections using two wing spars, one at the leading edge and the other at the two-third chord. The portion of the wing between the spars was flexible and the portion aft of the second spar rigid. The wind tunnel test results showed that the wing had a maximum lift coefficient of 0.76 and minimum drag of 0.007. Sachs and Mehlhorn (1991, 1993) presented a method for endurance maximization, which consisted of the periodic optimal control of camber in a coordinated process with a corresponding control of throttle and elevator. It was shown that using camber control provides an efficient means to improve the L/D ratio at each flight condition during the unsteady phases of periodic optimal endurance cruise. Fuel consumption during the idle phase is also minimized. McGowan et al. (1999, 2008) outlined the morphing wing research activities at NASA Langley and also presented wing concepts that allowed shape changes

AIRFOIL ADJUSTMENT Airfoil adjustment is mainly concerned with camber variation, although there is also some research concerned with thickness change. Tables 911 summarize the literature available, with an additional categorization using the adopted activation method/strategy. In aerodynamics, camber represents the effective curvature (or shape) of an airfoil. The term camber control simply refers to the change of the curvature of the airfoil by means of actuators. There are many methods that can be implemented in order to effectively vary the camber. The wing camber can change either on specific parts (leading or trailing edge) or in a global manner, letting the entire wing act as a unique control surface. The choice of actuators can be conventional (i.e., electromagnetic motors, hydraulic, pneumatic, and actuators with moving parts) or solid-state smart materials (i.e., piezoelectrics, SMAs, rubber muscle actuators (RMAs), magnetostrictive materials, etc.). The actuation (whether it is induced

850

Table 9. Airfoil morphing: Camber with conventional actuation.


Vehicle Experimental Category Fixed wing Fixed wing Fighter Control and performance (L/D, CLmax)     Motor Electro-hydraulic Rigid Flexible   Size Actuation Skin Purpose Numerical Prot. WT FTB 8  Investigation

References Camber Camber

Geometrical parameters

Boeing (1973) de Camp and Hardy (1984) Bonnema and Smith (1988) Smith and Nelson (1990) Smith et al. (1992) Powers et al. (1992) Szodruch and Hilbig (1988) Camber 8   Performance (L/D, CLmax) Performance (CL, CLmax)      Flexible Stretchable Flexible Stretchable Flexible UAV Transport MAV Transport Servo Flexible Rigid Flexible Rigid  8     8  UAV Servo Servo Rigid Membrane Flexible Performance (L/D)    Camber (TE) Camber (TE) Camber Camber and twist Camber (TE) Camber Camber and twist Servo Pneumatic Servo Active Camber Camber Camber (TE) Camber Camber and twist Camber Camber Camber (TE) Camber (TE) Camber (TE) Camber (TE) Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing Fixed wing MAV Fixed wing Transport Fixed wing Fixed wing Transport UAV Fixed wing Fixed wing MAV Fighter UAV Passive and servo Ultrasonic Motor Membrane Flexible Pneumatic Flexible Fixed wing Transport Hydraulic Rigid Fixed wing Linear-wave motor Rigid Fixed wing Transport Rigid 8 8         8 8 8 8  8 8 8 8 Performance (L/D) Performance (CD, Range)

 8 8 8       8 8 8 8  8 8 8 8

8 8 8 8  8   8  8 8 8 8  8 8 8 8 (continued)

Austin et al. (1997)

Monner (2001) Monner et al. (1998) Campanile (2008) Campanile and Sachau (2000) Ifju et al. (2001) Bartley-Cho et al. (2004)

S. BARBARINO ET AL.

van Dam (2002) Garcia et al. (2003) Boothe (2004) Poonsong (2004)

Shkarayev et al. (2004)

Performance (CL, CLmax) Performance (L/D) Performance (CL, CLmax)

Baker et al. (2005) Baker and Friswell (2009) Bornengo et al. (2005)

Gern et al. (2005) Lajux and Fielding (2005)

Null and Shkarayev (2005)

Edi and Fielding (2006)

Performance (L/D) Control (roll) Performance (CL) Performance (CL, CLmax) Performance (CL, range)

Santangelo et al. (2006) Stanford et al. (2007)

Diaconu et al. (2008)

Table 9. Continued.
Vehicle Investigation Experimental Category Fixed wing Rotary wing Fixed wing Transport Servo and hydraulic Motor Flexible Flexible Size Actuation Skin Purpose Numerical    Prot. 8   WT 8 8 8 FTB 8 8 8 8 8 8 Performance (L/D) Flexible   Flexible Rigid Servo Flexible      8 8    8 8 8 8  8 8 8

References Camber Camber (LE) Camber (TE) Camber (TE) Camber (TE) Camber Fixed wing Fixed wing Servo Electromagnetic Rotary wing Fixed wing Fixed wing Fixed wing Fixed wing Small UAV UAV Transport Transport MAV Camber (TE) Camber (TE) Camber (LE) Camber (TE) Camber (TE)

Geometrical parameters

Frank et al. (2008)

Kota et al. (2008)

Performance (L/D) Performance (CLmax)

Morphing Aircraft

Ricci (2008) Miller et al. (2010) Shili et al. (2008) Wildschek et al. (2008) Boria et al. (2009)

Flexible Flexible Flexible

  

  

8 8 

Daynes et al. (2009) Kota et al. (2009)

Monner et al. (2009)

Marques et al. (2009)

Performance (L/D, range) Performance (CL, CLmax) Performance (CD)

Perera and Guo (2009)

851

852

S. BARBARINO ET AL.

Table 10. Airfoil morphing: Camber with SMA (and similar) actuation.
Vehicle Geometrical param. Camber (TE) Camber Investigation Experimental Authors Roglin and Hanagud (1996) Roglin et al. (1996) Kudva et al. (1996b) Martin et al. (1999) Florance et al. (2003) Kudva (2004) Sanders et al. (2004) Beauchamp and Nedderman (2001) Chopra (2001) Epps and Chopra (2001) Singh and Chopra (2002) Strelec et al. (2003) Alasty et al. (2004) Benavides and Correa (2004) Madden et al. (2004) Yang et al. (2006) Mirone (2007) Mirone and Pellegrino (2009) Song and Ma (2007) Brailovski et al. (2008) Peel et al. (2009) Chou and Philen (2008) Popov et al. (2008) Seow et al. (2008) Abdullah et al. (2009) Barbarino et al. (2007) Icardi and Ferraro (2009) Lv et al. (2009) Barbarino et al. (2010a) Category Rotary wing Size UAV Actuation SMA Rigid Skin Purpose Performance (CLmax) Num.   Prot.   WT   FTB  8

Fixed wing

Fighter UAV

SMA

Flexible

Camber Camber (TE)

Rotary wing Rotary wing

Wind mill

SMA SMA

Flexible

Performance ( C L)

8 

8 

8 

8 8

Camber Camber (TE) Camber (TE) Camber (TE) Camber Camber

Fixed wing Fixed wing Fixed wing Rotary wing Fixed wing Fixed wing Marine Propeller Small UAV UAV

SMA SMA SMA PPy SMA SMA

Flexible Flexible Flexible Rigid Flexible Flexible

Performance (L/D) Performance (L/D, CLmax) Performance (L/D) Performance ( C L) Performance (L/D)

    8  

 8        8 8  8

 8  8  8

8 8 8 8 8 8

Camber (TE) Thickness Camber Camber (TE) Thickness Camber (TE) Camber and thickness Thickness Camber Camber (TE) Camber and twist Camber (TE)

Fixed wing Fixed wing Fixed wing

UAV

SMA SMA RMA FMC SMA Flexible Flexible Flexible Flexible Stretchable Flexible Performance (L/D) Performance (CD)

8 8 8 8 8 8 8

8 8 8 8 8 8 8

        8

Fixed wing Fixed wing Small UAV UAV

SMA SMA

Fixed wing Fixed wing Fixed and rotary wing Fixed wing

Transport UAV

SMA SMA SMA

Flexible Corrugated and stretchable

 8  8

8 8 8 8

8 8 8 8

Transport

SMA

Stretchable

Morphing Aircraft Table 11. Airfoil morphing: Camber with piezoelectric (and similar) actuation.
Vehicle Geometrical parameters Camber (TE) Camber and twist Camber (LE) Camber Investigation

853

Experimental Authors Spangler and Hall (1990) Lazarus et al. (1991) Pinkerton and Moses (1997) Lee (1999) Chopra (2001) Koratkar and Chopra (2000) Lee and Chopra (2001a, b) Bernhard and Chopra (2001) Geissler et al. (2000) Munday and Jacob (2001) Straub et al. (2001, 2009) Wang et al. (2001) Barrett (2002) Barrett and Tiso (2004) Barrett et al. (2005) (XQ-138) VT Senior Design, Eggleston et al. (2002) Vos et al. (2007, 2008a) Category Rotary wing Size Actuation PZT Skin Rigid Flexible Fixed wing Rotary wing PZT PZT and magnetostrictive Membrane Purpose Control Performance (CL, Weight) Performance (L/D, CLmax) Control and performance Numerical   8  Prot.  8   WT  8   FTB 8 8 8 8

Camber (LE)

Camber and thickness Camber (TE) Camber and twist Camber

Fixed and rotary Fixed wing Rotary wing Fixed wing Rotary wing

PZT

Rigid

PZT PZT Fighter UAV UAV PZT and others PZT

Membrane Flexible Flexible Flexible and membrane

Performance (L/D, CLmax)

8  

   

   

8 8 8 

Performance and control

Camber

Fixed wing Rotary and fixed wing Fixed wing Rotary wing Rotary and fixed wing Fixed wing Fixed wing Fixed wing

Small UAV Small UAV

PZT, SMA, servo PZT

Flexible

 

 

 

Camber

Membrane

Bilgen et al. (2007, 2009) Grohmann et al. (2006, 2008) Bilgen et al. (2009, 2011a)

Camber (TE) Camber (TE) Camber

Small UAV

PZT PZT

Flexible Flexible Flexible

Control (roll)

  

  

 8 

 8 8

Small UAV

PZT

Paradies and Ciresa (2009) Wickramasinghe et al. (2009) VT Senior Design, Butt et al. (2010a), Bilgen et al. (2011b)

Camber Camber (TE) Camber (TE)

UAV UAV Small UAV

PZT PZT PZT

Flexible Membrane Flexible

Performance (L/D, CLmax)

 

  

 8 

8 8 

Control (roll, pitch, and yaw)

854

S. BARBARINO ET AL.

without surface discontinuities. In one concept, referred to as the fish bone concept, the main load-bearing component resembles the spinal cord of a fish covered with an elastomeric material to transfer pressure loads to the main structure. The trailing-edge ribs can deflect up to 20 in camber and 25 in the span-wise direction. Likewise, the leading-edge ribs can deflect 25 in camber and 20 in the span-wise direction. The configuration was built for the assessment and demonstration of the structural concept and not for wind tunnel testing. Bolonkin and Gilyard (1999) calculated the benefits of variable-camber ailerons that span the majority of the wing. The main purpose is maximizing the L/D ratio and reducing the fuel consumption of subsonic transport aircraft. Recent attention is given to multi-objective and multilevel (or multiscale) optimization problems due to the increasing computational power. The design of morphing aircraft (or airfoil) geometries is now commonly posed as a multilevel, multi-objective optimization problem. In most morphing aircraft problems, the two competing objectives are (1) maneuverability and (2) long range/endurance. In general, multi-objective problems have many optimal solutions each depicting a different compromise scenario. Each optimal solution is known as a Pareto point, and the set of all these points represents the Pareto curve. The Pareto description is known to be a powerful means of showing the global picture of the solution field. Prock et al. (2002) explored a process to link analytical models and optimization tools to create energy-efficient, lightweight wing/structure/actuator combinations for morphing aircraft wings. The energy required to deform the airfoil was used as the performance index for optimization while the aerodynamic performance such as lift or drag was constrained. Three different, but related, topics were considered: energy required to operate articulated trailing-edge flaps and slats attached to flexible two-dimensional (2D) airfoils; optimal, minimum energy, articulated control deflections on wings to generate lift; and, deformable airfoils with cross-sectional shape changes requiring strain energy changes to move from one lift coefficient to another. Results indicate that a formal optimization scheme using minimum actuator energy as an objective and internal structural topology features as design variables can identify the best actuators and their most effective locations so that minimal energy is required to operate a morphing wing. Gano and Renaud (2002) presented a concept to increase the efficiency of a UAV, based on changing the airfoil thickness. They suggested to decrease the volume of the wing tanks as the fuel is depleted, thus inducing a thickness reduction and thereby a drag reduction. Gano et al. (2004) discussed solution strategies for multi-objective, multi-level morphing aircraft design problems. Two design tools were explored to combat the issues in the

optimization problem: conversion to a single-level design problem and the use of variable-fidelity optimization. A variable-fidelity optimization framework was discussed and applied to design a buckle wing morphing aircraft concept. The buckle wing in cruise conditions looks like a typical UAV with a standard high-aspect ratio wing; though the wing may be slightly thicker. However, the wing is actually composed of two thinner wings which are fused together to form this high-aspect ratio wing, and when the need for maneuverability arises the aircraft can, via a buckling load applied at the wing tips, morph into a configuration that resembles a biplane with its outboard wing tips joined together. When the UAV is in the buckled state, it can generate much more lift and becomes more agile. The aerodynamic models used in the study include a high-fidelity CFD model and a low-fidelity panel method. Rusnell et al. (2004) also investigated the buckle-wing concept. Compromise Programming was used as the optimization method and the Vortex-Panel Method was used to calculate the aerodynamic behavior. The buckle-wing UAVs enhanced capabilities were demonstrated both quantitatively and qualitatively. Johnston at al. (2003, 2007) analyzed the resistance of an airfoil to change its shape due to elastic and aerodynamic forces. The work required to overcome this force was examined in a variety of different conditions. The energies required for a pitching flat plate, conventional flap, conformal flap, and two variable-camber configurations were investigated. It was determined that the distribution and the size of actuators can make a significant difference in the energy required. Namgoong et al. (2006) presented the strain energy required to change the shape (camber and thickness distribution) of a wing section. The additional power input to the morphing section was taken as a new design constraint. This research determined the energy required to deform airfoil sections, and proposed that the morphing airfoil design problem is multi-objective when the aerodynamic constraints and strain energy problems are both solved. Zhao et al. (2009) presented a robust airfoil design problem. The authors reported that the performance of an initial airfoil is enhanced through reducing the standard deviation of CLmax. They also concluded that the maximum thickness has the dominant effect on the mean value of CLmax, the location of maximum thickness has the dominant effect on the standard deviation of CLmax, and the maximum camber has a little effect on both the mean value and the standard deviation. Camber Motivation for Adoption in Rotary Wings In addition to the fixed-wing applications above, rotary aircraft have been an active research area with the introduction of piezoelectric materials for vibration control. Although the references presented next do not

Morphing Aircraft

855

propose camber morphing, the concepts are still capable of camber control and they provide a good background to some of the camber morphing research today. In the early 1990s, Steadman et al. (1994) showed an application of a piezoceramic actuator for camber control in helicopter blades. Structurecontrol interaction was employed to develop an adaptive airfoil that can be used in the cyclic and vibration control of the helicopter. Giurgiutiu et al. (1994a, b) researched improved rotor blades using strain-induced actuation methods (using PZTs). Wind tunnel experiments proved the control authority of the PZT actuators. Giurgiutiu (2000) presented a comprehensive review on the application of smart material actuation to counteract aeroelastic and vibration effects in helicopters and fixed-wing aircraft. Experiments of active flutter control, buffet suppression, gust load alleviation, and sonic fatigue reduction were discussed. Kerho (2007) presented analysis to reduce the effects of dynamic stall on rotorcraft blades. The adaptive airfoil had an aerodynamically smooth variable leading-edge droop which enabled dynamic stall control. Such a concept incorporates the change in camber to effectively postpone the occurrence of stall. A modified NavierStokes solver (OVERFLOW) was employed and the results clearly showed that the compliant structure had a higher CL than a baseline section, or could eliminate the dynamic stall vortex at a CL equivalent to the baseline section CLmax, while maintaining the baseline sections high-Mach number advancing blade characteristics. Gandhi et al. (2008b) proposed a variablecamber airfoil as an alternative to trailing-edge flaps used for active helicopter vibration reduction. The design consists of compliant mechanisms actuated using PZTs, with the variable camber-achieved aft of the leading-edge spar. An optimized shape was predicted to achieve a trailing-edge deflection angle of 4.6 , resulting in a lift increase of 1722%. The aerodynamic loads were found to cause only small deformations in comparison with those caused by the actuation. Bench-top tests demonstrated that rotor airfoil camber is controllable using the proposed concept. In a complementary study, Gandhi and Anusonti-Inthra (2008a) focused on identifying favorable attributes for a skin for a variablecamber morphing wing. The paper proposed that the skin should be highly anisotropic, or more specifically (1) has a low in-plane axial stiffness (so it can expand and contract with the change in camber) and (2) has a high out-of-plane flexural stiffness (to resist deformation due to out-of-plane aerodynamic loading). Camber Motivation for Adoption in Small Aircraft The interest in MAVs and their morphing capabilities is connected with the recent and increasing development of UAVs. These aircrafts have similar flight conditions to insects and birds, as they fly at low Reynolds and

Mach numbers (Santhanakrishnan et al., 2005). MAVs represent a new challenge for aerodynamics, propulsion, and control design. Typical MAV missions are projected to be flown either by inexperienced operators or via autonomous control. This requires robust flying characteristics. It is well known that during flight in the Reynolds number range between 10,000 and 100,000, flow separation around an airfoil can lead to sudden increases in drag and loss of efficiency. The effects of flow separation can be seen in nature where large species soar for extended periods of time while small birds have to flap vigorously to remain airborne. The Reynolds numbers of the larger species are well above 100,000, whereas hummingbirds fly at below 10,000 if they attempt to soar. Aerodynamic efficiency is critical in MAV development, and the low-Reynolds number regime is relatively new to aeronautical applications. There are several benefits of employing continuous shape (or more specifically camber) control via active materials over discrete trailing-edge control using conventional control surfaces in small air vehicles. First, the low-Reynolds number flow regime can result in flow separation that reduces the effectiveness of a trailingedge control surface. Second, small UAVs and MAVs cannot afford to lose energy through control surface drag because of their severe power limitations. Finally, the opportunity for flow control is inherent in the active material due to its direct effect on circulation and its high operating bandwidth. In addition to replacing conventional control surfaces for camber control, these actuators can be effective in dynamic laminar separation bubble control (also referred to as flow control) although this research field is not considered here. Shyy et al. (1997, 2010) presented research on airfoils whose camber changes in response to aerodynamic loads, including a review of recent progress in flapping wing aerodynamics and aeroelasticity. Instead of actively morphing the wing section, the airfoils described in the review were deformed passively. Flexibility in the airfoil section was shown to improve unsteady airfoil performance by limiting flow separation at high angles of attack. However, for moderate angles of attack where flow separation on the rigid airfoil is limited, the flexible airfoil suffers from a slightly degraded lift-to-drag ratio compared to the rigid case. The computational study was followed by an experimental study in a low-speed wind tunnel with an unsteady flow velocity, where flexible membrane and rigid flat airfoil sections were tested, as well as a hybrid section composed of a membrane reinforced with stiff wires. Passive morphing appears attainable for small MAVs operating in the Reynolds number range where thin plate airfoils can be effective. While no benefits are realized for steady flow conditions, flexible camber membranes do appear to offer benefits for vehicles such as helicopters and ornithopters. Song et al. (2001) focused on the measurement of the upward

856

S. BARBARINO ET AL.

and downward strokes of a dragonfly wing. The paper analyzed the camber change via optical fringe patterns. The experimental results showed that the camber deformation was significantly different during the upstroke and the downstroke and this difference is fundamental to the generation of lift. Barrett (2004) reviewed several families of adaptive aerostructures and flightworthy aircraft, including rotary and fixed wing vehicles, munitions, and missiles. More than 40 adaptive aerostructure programs which have had a direct connection to flight test and/or production UAVs, ranging from hover through hypersonic and from sea-level to exo-stratospheric, were examined. A historical analysis showed the evolution of flightworthy adaptive aircraft from the earliest flights in 1994 to the publication date in 2004. These aircrafts are discussed in different sections of current review paper. Mesaric and Kosel (2004) investigated the prediction of unsteady airloads imposed on a variable-camber wing that change over time with deformation. Several methods were used for the analysis. Results from an analytical method based on incompressible, potential flow theory were validated via a numerical time-step method. This analysis also had important vibration and flutter implications. Murua et al. (2010) numerically investigated the effect of chord-wise flexibility on the dynamic stability of compliant airfoils. A classical twodimensional aeroelastic model was expanded with an additional degree of freedom to capture the time-varying camber deformations. A number of situations were identified in which the flutter boundary of the compliant airfoil exhibited a significant dip with respect to the rigid airfoil models. The results can be used to estimate the aeroelastic stability boundaries of membrane-wing MAVs. Kim and Han (2006) and Kim et al. (2009) designed and fabricated a smart flapping wing using a graphite/ epoxy composite material and a macro-fiber composite (MFC) actuator. This research aimed to mimic the flapping motion of birds. The wind tunnel tests were performed to measure the aerodynamic characteristic and performance of the surface actuators. The test vehicle was the Cybrid-P2 commercial ornithopter with and without the MFC actuator. A 20% increase in lift was achieved by changing the camber of the wing at different stages of flapping motion. Hu et al. (2008) employed particle image velocimetry to gain an in-depth understanding of the underlying aerodynamics in membrane wings. The rigidity of the membrane wing was varied by changing the number of evenly spaced ribs throughout the wing surface. The trailing edge of the airfoil was observed to deflect with increasing load, thus reducing the effective angle of attack of the entire section and delaying the onset of stall. The aerodynamic efficiency of flexible airfoils was also confirmed in the study,

demonstrating an increase in L/D for more compliant wings, except when the increased compliance induces trailing-edge vibrations and/or instabilities. Song et al. (2008), inspired by the membrane wings of small flying mammals, performed a broad study of the effects of membrane geometry and elasticity on aerodynamic performance. In this case, the wing was a sheet of latex stretched between two rods at various aspect ratios, and rigidity was increased by separating the rods further from each other, stretching the material further. The wings compliance to the airflow results in an increase in aerodynamic efficiency, suggesting that the wing adapts itself to a variety of flight regimes without any cost in energy to whatever system may employ it. In other words, instead of using actuators to change the airfoil geometry and maximize lift, a membrane wing does so automatically. Increasing compliance generally leads to higher cambers at lower aerodynamic loads, but can operate at higher angles of attack and stalls more gently than less compliant or rigid wings. Drag also increases with compliance, but an improvement in L/D is still obtained compared to rigid wings. Turnock and Keane (2009) considered several approaches to minimize actuation requirements of aerial and maritime autonomous vehicles. A concept uses a combination of pushpull actuators coupled with a snap-through composite lay-up to achieve shape change. It was proposed that such a system could be applied to the trailing edge of an autonomous underwater glider wing. The anisotropy achieved through use of different composite ply orientations and stacking sequence could also be used to generate bendtwist coupling such that fluid dynamic loads induce passive shape adaptation. Another approach used a detailed understanding of the structural response of buckled elements to apply control moments to deform a complete wing surface. Conventional Actuation for Camber The most common form of camber variation uses conventional control surfaces. Camber control is widely used in modern aircrafts in the form of elevators, rudders, ailerons, flaps, etc. By simply deflecting a control surface on the airfoil, the camber of the airfoil is effectively increased along with the angle of attack, increasing the lift across that particular surface. Trailing-edge control is the more popular of the two, although the F16 Fighting Falcon uses leading-edge flaps to change the wing camber, for example. All commercial aircrafts today utilize discrete control surfaces that are actuated via conventional devices. Conventional systems are typically lumped systems with electromagnetic, hydraulic, pneumatic or motor (with rotor and stator parts) actuators, and a revolute or a prismatic joint system(s). The designs that are currently in use (in the aerospace

Morphing Aircraft

857

industry) are discrete and rotating leading- and trailingedge controls. While some of those mentioned (earlier) may already be in use (i.e., leading electrolyte (LE) and trailing electrolyte (TE) devices), those that focus on morphing the complete airfoil to provide control of the aircraft are still in development and are more interesting. Complete camber morphing offers huge potential for advancement in the way modern and future aircraft perform; however, due to its complexity, this method is often utilized in small unmanned aircraft. In this section, attention is given to camber morphing for fixed-wing and rotary aircraft in the subsonic regime to achieve a smooth variation of the airfoil shape using lumped (not distributed) actuator systems. Boeing (1973) presented an advanced technology variable-camber wing wind tunnel test in the NASA Ames 14-ft transonic wind tunnel. The wing had simple hinged leading- and trailing-edge flaps but also smooth and curved variable-camber flaps. Large improvements over standard airfoil performance were obtained, but the device was mechanically complex in the wing interior. Starting from the mid-80s, the researchers at Boeing (de Camp and Hardy, 1984; Bonnema and Smith, 1988; Smith and Nelson, 1990) integrated, on board a military aircraft, a device capable of controlling the wing curvature by means of an automated control system. The goal was to optimize performance according to the external wing pressure loads. This program, named Advanced Fighter Technology Integration (AFTI)/F-111, was promoted by NASA in collaboration with the USAF. The F-111 wing was suitably modified, adopting six independent sections for the trailing edge (three for each wing, based on sliding panels for the lower surface and flexible panels made of glass fiber for the upper one) and two for the leading edge in flexible composite; all these surfaces were commanded by electro-hydraulic actuators. Flight tests were conducted on the AFTI/F-111 aircraft and confirmed the performance increase: 2030% range enhancement, 20% aerodynamic efficiency growth, and 15% increase of wing airload at a constant bending moment. Bonnema and Smith (1988) presented the comparative analysis between Mission Adaptive wings and aircrafts with fixed wings under the research program. Smith et al. (1992) focused mainly on integration of the variablecamber control into the aircraft and assessed the performance and reliability. Powers et al. (1992) presented additional flight test results from the Mission Adaptive Wing program. Szodruch and Hilbig (1988) presented a variablecamber wing concept to reduce drag and hence fuel consumption of a transport aircraft. The research gave various wind tunnel results on wings of different cambers, wing design philosophy, systems requirements, and finally the benefits of new wing concept. Austin et al. (1997) proposed an adaptive trailing edge actuated by a

linear-wave motor with magnetostrictive material (Terfenol-D), which provided wing morphing through the internal structure displacement. Monner et al. (1998) and Monner (2001), within the ADIF project carried out by EADS-Airbus, DaimlerChrysler F&T and DLR, developed a concept to replace rigid, fixed ribs with a flexible alternative with a high stiffness (intended for large transport aircraft). The flexible plates were realized by combining several plates, contoured to the airfoil section shape, with revolute joints. A single actuator was then used to actuate the first plate in the series, which transferred the motion to the adjoining plates via the kinematic constraints imposed by the joints. In this way, the desired degree of camber was achieved. The skin is allowed to move relative to the joint by incorporating slide joints and stringers to maintain stiffness. A proof of concept achieved the desired degree of camber. The research concentrated on the optimization of the rib elements to achieve minimum joint stress and maintain the camber objectives. The actuation was hydraulic and the concept was considered as lift augmentation/drag reduction rather than direct control. Santangelo et al. (2006) developed and simulated a servo-actuated articulated rib, similar to the original work presented by Monner et al. (1999). Monner et al. (2009) focused on leadingedge morphing, based on a collaboration between DLR and EADS within the SmartLED project. A smart leading-edge device was developed, which delivered an alternative to the droop nose device used by the A380: it was developed from Dornier Patent DE 2907912 using a suitable optimization process. The main emphasis of this new device was to realize a structure/system solution for a smooth leading surface, which can be deflected in a typical high-lift application: a deflection of 20 was selected as the target. Campanile and Sachau (2000) and Campanile (2008) described a method of airfoil shape morphing using structronic instead of mechatronic mechanisms. The concept is to replace the rigid rib in an airfoil with a belt rib, which is a compliant structure that can be realized with a conventional material. The basic form of the airfoil was achieved using in-plane stiff spokes attached to the belt-rib by solid-state hinges. Due to the relatively low bending stiffness of the hinges, the spokes are mainly loaded in a tensioncompression mode. During the deformation in the shape, these spokes can be assumed rigid. With this assumption, the shape behavior of the belt-rib structural frame is defined by the belts bending flexibility and the configuration of the spokes. The variation in camber was achieved by several methods such as conventional actuation or embedded smart material actuation. Static (proof of concept) bench tests were conducted using mechanical wires and SMA actuation although no details of the actuation suitable for specific aircraft applications were given.

858

S. BARBARINO ET AL.

Bartley-Cho et al. (2004) presented the smooth variablecamber wing that was applied in a Northrop Grumman UCAV test model (part of the Smart Wing Program, discussed in the next section). The unmanned aircraft demonstrated high actuation rate (80 /s), large deflection (20 ), hinge-less, smoothly contoured control surfaces with chord-wise and span-wise shape variabilities. Ultrasonic (piezoelectric material driven) motors were used as actuators. van Dam (2002) presented a comprehensive review of recent developments in aerodynamic design and analysis methods for multi-element high-lift systems on transport airplanes. Attention was given to the associated mechanical and cost problems since a multi-element high-lift system must be as simple and economical as possible while meeting the required aerodynamic performance levels. Poonsong (2004) designed and implemented a multisection wing model based on the NACA 0012 airfoil, capable of variable curvature. Each rib was divided into six sections, each able to rotate up to 5 with respect to the previous one without significant discontinuities on the wing surface. The camber change was obtained by means of pneumatic actuators. Wind tunnel tests showed that the lift produced by a similar wing was comparable to that of the same one piece airfoil, but the surface drag was higher due to the increased required flexibility of the skin. Baker et al. (2005) and Baker and Friswell (2006, 2009) proposed a compliant structure to effect a change in airfoil camber. The trailing edge of the airfoil section was morphed using a combination of 3 or 88 actuators placed within 14- or 1752-element truss structures, respectively. The trusses supported and defined the shape of the airfoil. The focus of the paper was to select which truss elements to replace by active materials capable of extending or contracting their length while minimizing the required actuation strain. Several optimization algorithms were evaluated the best optimization being performed by a relatively timeconsuming genetic algorithm (GA). This sort of concept is well suited to actuation by SMAs or piezoelectric stacks. Bornengo et al. (2005) examined a possible internal wing structure that is adaptive to shape/camber changes. The chiral honeycomb design differed from classical wing box structures and presented unique structural behavior in terms of its effective Poissons ratio. The paper details the numerical finite element analysis models used to better predict a homogenized material behavior. It was determined that the material would change camber with changes in flow speed. Gern et al. (2005) proposed the replacement of trailing-edge flaps by distributed structural actuation. The power requirements for the actuation of the conventional flaps were compared to that of the morphing wing design. Lajux and Fielding (2005) developed a new methodology for wing leading-edge devices for variable camber

applications. The research provided insight into CFD simulation, new mechanism design solutions, and structural analysis. It also described estimation methods for weight and reliability and trade-off studies of alternative configurations. The methodology proposed aimed to speed up the design process while incorporating the variable camber technology. A comparative study between conventional LE mechanisms and variable camber technology was also given. Edi and Fielding (2006) examined the viability of using variable camber airfoils on transport aircraft. An airfoil employing hybrid laminar-flow control (at the LE) and variable camber (at the TE) was considered and a three-dimensional CFD analysis used to determine the aerodynamic properties of the design. Diaconu et al. (2008) presented a non-linear static finite element study of a morphing airfoil to determine the forces required for various levels of deflection for a graphite/epoxy composite material. The study proposed three bi-stable concepts and the respective geometric properties, including a control surface fixed between two vertical spars with a bi-stable adaptive trailing edge. Flexsys Inc. (2011; Website) focused on the aerodynamic smoothness of the wing and developed a functional, seamless, hinge-free wing whose trailing and/or leading edges morphed on demand to adapt to different flight conditions. Kota et al. (2009) presented results (from Flexsys) on the variable-camber trailing edge for a high-altitude, long-endurance aircraft. The flight testing of the Mission Adaptive Compliant Wing adaptive structure trailing-edge flap used in conjunction with a natural laminar flow airfoil was described. The airfoil flap system was optimized to maximize the laminar boundary layer extent over a broad lift coefficient range for endurance aircraft applications. The wing was flight tested (in 2006) at full-scale dynamic pressure, full-scale Mach number, and reduced-scale Reynolds numbers on the Scaled Composites White Knight aircraft. Data from flight testing revealed that laminar flow was maintained over approximately 60% of the airfoil chord for much of the lift range. Drag results were provided based on a dynamic pressure scaling factor to account for White Knight fuselage and wing interference effects. The expanded laminar bucket capability allows the endurance aircraft to significantly extend its range (15% or more) by continuously optimizing the wing L/D throughout the mission. Frank et al. (2008) mechanized a controlled spanwisevarying airfoil camber change for a high-aspect ratio wing, resulting in optimized aerodynamic performance for an aircraft that changed weight by 50% over its mission. Two separate design methodologies were used to achieve these shape changes: (1) a rigid body kinematics approach and (2) a compliant mechanism approach. A framework to optimize these two mechanisms was presented. The designed mechanisms were evaluated based

Morphing Aircraft

859

on the error in the shapes and on the energy efficiency of the systems. Ricci (2008) developed a rotating rib concept capable of gapless camber variation, achieved by an internal hinge to which the trailing-edge rib was attached. Both the upper and lower skins were allowed to glide over the rib contour (they were not connected at the trailing edge). A wing section was fabricated and equipped with four rotating ribs driven by four servo actuators and a maximum rotation of 5.5 was demonstrated. Miller et al. (2010) presented the research of the SMorph (Smart Aircraft Morphing Technologies) project which was a Eurocores S3T collaboration. The aim of the project was to develop novel structural concepts to implement morphing aeroelastic structures. An overview of the current status of the project was provided, including descriptions of the development, modeling, and design optimization of several morphing aeroelastic structures. The design, fabrication, and bench-testing of concepts were presented. Shili et al. (2008) introduced a systematic approach to design compliant structures to perform required shape changes under distributed pressure loads. The distributed compliant mechanism was optimized using a GA. A direct search method was used to locally optimize the dimension and input displacement after the GA optimization. The resultant structure achieved a 9.3 angle change, which was also validated using a prototype. Wildschek et al. (2008) proposed an all-composite morphing trailing-edge design for flying wing aircraft control. An adaptive wing demonstrator with an allelectric actuation system was proposed by splitting the trailing edge of the wing into upper and lower seamless control surfaces. The upper and lower structures were moved independently by two electric actuators. These inner structures of both upper and lower surfaces were hinged to several bars which were rigidly connected to the upper and lower surfaces. This multi-functional morphing trailing edge was shown to provide full control authority, such as roll, pitch, yaw, and load control, high lift, and airbrake function at the same time. The proposed morphing trailing edge device achieved a maximum deflection angle of 35 . In addition to the fixed-wing applications (presented above), researchers at FlexSys and Kota et al. (2008) addressed the problem of helicopter retreating blade stall through the use of a morphing variable-camber airfoil capable of being actuated once per rotor revolution. Morphing of the airfoil camber was accomplished by a compliant structure that precisely deformed the leading edge of the airfoil section using a single linear electromagnetic actuator. Aerodynamic optimization was employed to identify a deformed airfoil shape that would delay flow separation at high angles of attack with minimal drag penalty. Likewise, structural optimization was used to find a structure capable of morphing

between the nominal and deformed airfoil shapes with low actuation force requirements, constrained by fatigue life limits. A model scale rotor was designed using a similar process, except that a rotary actuator was selected; the model was fabricated and shown to achieve a 6 Hz response rate. Daynes et al. (2009) presented the design and wind tunnel testing results of a full-scale helicopter rotor blade section with an electromagnetic-actuated bistable trailing-edge flap, developed under the Intelligent Responsive Composite Structures program. The flap system was designed to change between two stable positions when the helicopter moves between hover and forward flight conditions. The bistability of the flap system meant that the electromagnetic actuator was only required to transit between these two stable states, and not to maintain the states. The use of bistable composites as aerodynamic surfaces was proven to be possible, and stable trailing-edge deflections of 0 and 10 were achieved. The wind tunnel model clearly demonstrated a significant change in lift coefficient compared to the standard airfoil section when the flap was deflected. In the area of small UAVs and MAVs, Ifju et al. (2001) and the researchers at the University of Florida (Waszak et al., 2001; Ifju et al., 2002; Torres, 2002) developed a series of vehicles that incorporated a unique, thin, reflexed, flexible wing design. The wings were constructed of a carbon fiber skeleton and a thin flexible latex membrane. The flexible wing design reduced the adverse effects of gusts and unsteady aerodynamics, exhibited desirable flight stability, and enhanced structural durability. There have been numerous experimental and analytical studies on flexible-wing MAVs. Albertani et al. (2004) investigated the effects of a propeller on the aerodynamic characteristics of MAVs and the coupling with the wing flexibility in steady conditions. The wind tunnel test data provided a detailed account of the aerodynamics as they relate to the propeller effects and to the structural deformations, namely the wing flexibility. Garcia et al. (2003, 2005) and Boothe (2004) developed a fleet of small UAVs, with a carbon fiber fuselage and wings of thin flexible sheets. The wing camber change and twisting were obtained by means of torsion bars (one for each wing), implemented by servos inside the fuselage. The attained roll rate was much higher than the yaw rate, indicating that pure roll control through wing twist was almost possible. These model aircrafts show a high agility and maneuverability, which makes them hard to manage remotely. In a similar study, Stanford et al. (2007) implemented morphing (in the form of asymmetric twisting) through the use of a torque-actuated wing structure with thousands of discrete design permutations. A static aeroelastic model of the MAVs was developed and validated to optimize the performance of the torque-actuated wing structure.

860

S. BARBARINO ET AL.

Objective functions included the steady-state roll rate and the lift-to-drag ratio during such a maneuver. An optimized design was obtained through the use of a GA presenting significant improvements in both performance metrics compared with the baseline design. Boria et al. (2009) outlined an effective experimental technique for shape management optimization of a morphing airfoil: a GA with wind-tunnel hardware (strain-gage sting balance) in the loop. Optimal wing shapes were found to maximize the measured lift or efficiency for a range of angles of attack by allowing the wing shape control to be defined through deformation at a single actuation point (camber control) or two actuation points (camber and reflex control). Shkarayev et al. (2004) presented a method to alter the camber of an MAV using micro-servos. A servomotor was placed at the inflection point of the airfoil, rather than just deflecting the trailing edge of the airfoil. One issue was the increase in nose-down pitching moment, which was compensated by adding reflex to the wing. The method transforms the shape of the wing, rather than just deflecting the trailing edge. Null and Shkarayev (2005) presented four MAV wind-tunnel models with 3%, 6%, 9%, and 12% cambered thin airfoils. Aerodynamic coefficients were measured using a low-speed wind tunnel. Large positive, nose-up pitching moment coefficients were found with all cambers at the lowest Reynolds number. These results were verified by flight tests of MAVs utilizing these airfoils. The 3% camber wing gave the best lift-to-drag ratio of the four camber values and theoretically would be the optimal choice for high-speed, efficient flight. In contrast, the 6% and 9% camber wings were predicted to give the best low-speed performance because of their high liftto-drag ratios and mild pitching moments near their stall angles of attack. Marques et al. (2009) described a process by which the camber of flaps could be changed via a hinged rigid flap and presented the aerodynamic optimization of the flap parameters. Variable camber was proposed to improve aerodynamic efficiency in a UAV. Perera and Guo (2009) investigated the optimal design of a seamless aeroelastic wing (SAW) structure for a lightweight aircraft. A hinge-less flexible trailing edge control surface and an eccentuator actuation mechanism were proposed. An innovative sliding trailing edge was created by improving an existing curved torque beam design. The SAW camber was varied in a desirable shape with minimum control power demand (an open trailing edge was used for this purpose). This design concept was simulated numerically and demonstrated by a test model. Subsequently, the wing structure configuration was optimized and aeroelastic tailored. The results showed that the initial baseline weight can be reduced by approximately 30% under the strength criteria. The resulting reductions in the wingbox stiffness and aeroelastic

stability have been improved by applying aeroelastic tailoring. SMA and Similar Actuation for Camber In recent years, the development of innovative materials allowed the adoption of new techniques for wing morphing, enabling the substitution of traditional servo, pneumatic, hydraulic, or motor-like actuation mechanisms with more integrated and lighter actuation elements made of solid-state smart materials. The possibility of changing the airfoil shape by means of muscle-like actuators (for example, SMAs, RMAs, and other similar actuators that can contract with the application of thermal, electrical, and mechanical energies) is very attractive. By directly implementing these materials within the structural elements, an actuation capability is integrated within the structure, with significant benefits in terms of: (1) weight, (2) reliability and maintenance, and (3) structural and aerodynamic efficiencies. The use of smart material systems can allow for both load-bearing and deformable structures, with the possibility of implementing a continuous, variable, adaptive, and integrated actuation. Roglin et al. (1994) and Roglin and Hanagud (1996) presented an adaptive airfoil that used a SMA actuator mechanism to actively change the camber of an airfoil for a remotely piloted helicopter. The airfoil concept was demonstrated on a remote control helicopter with active camber airfoils. Embedded SMA wires were used to bend the trailing edge of the rotor blade to morph the airfoil camber. This camber morphing was then used to replace the collective pitch control of the helicopter. The change in rotor thrust response due to the SMA is linearized by a controller which accepts the signal from the radio receiver and applies the heating required to deform the SMA. Due to the low-frequency response of the system, only steady collective control was attempted. Kudva et al. (1996a, b), Kudva and Carpenter (2000), Kudva (2001), and Florance et al. (2003) worked on SMA applications for a morphing wing. Kudva (2004) documented the DARPA, AFRL, NASA, Northrop Grumman Smart Wing program as it evolved since its inception in 1995. The program aimed to realize a UCAV and considered the application of smart materials to improve the performance of military aircraft. The project consisted of two phases and both phases had experiments with SMAs. The program utilized an SMA to contour the trailing-edge control surfaces, and SMA torque tubes to vary the wing twist. Early efforts highlighted the bandwidth issues of SMAs. During the second phase, a 30% scaled design was tested over a range of Mach 0.30.8, and resulted in significance performance improvements (increased lift, and a roll rate of 80 /s). From wind tunnel tests, an improvement between 8% and 12% in lift and roll control over a traditional

Morphing Aircraft

861

wing was observed. A wing twist up to 5 was attained, together with a deflection of up to 4.5 for the leading edge and 15 for the trailing edge. Kudva (2004) mentioned that the Smart Wings control surfaces showed no degradation in performance during the 3 weeks of wind tunnel testing. Martin et al. (1999, 2002) presented the design and fabrication details of the two Smart Wing Phase 1 wind tunnel models. Sanders et al. (2004) presented the wind tunnel results obtained during Phases 1 and 2 of the Smart Wing program. Chopra (2001, 2002) and Epps and Chopra (2001) systematically investigated the development of an SMA-actuated trailing-edge tab for in-flight blade tracking. This wing section was tested in the open-jet wind tunnel, and tab deflections in the order of 20 were obtained. Singh and Chopra (2002) improved this design and successfully tested it in the wind tunnel with repeatable open-loop and closed-loop performances. Beauchamp and Nedderman (2001) patented a design with SMAs that were embedded within a control surface on the blade of a wind turbine. Strelec et al. (2003) utilized a global optimization method that incorporated a coupled structural, thermal, and aerodynamic analyses to determine the necessary placement of the SMA wire actuators within a compliant wing. A GA was chosen as the optimization tool to efficiently converge to a design solution. The GA used was a hybrid version with global search and optimization capabilities augmented by the simplex method with selective line search as a local search technique. The liftto-drag ratio was maximized for a reconfigured airfoil shape at subsonic flow conditions. A wind tunnel model of the reconfigurable wing was fabricated based on the design optimization to verify the predicted structural and aerodynamic responses. Wind tunnel tests indicated an increase in lift for a given flow velocity and angle of attack by activating the SMA wire. Alasty et al. (2004) studied the effect of a variableshape wing applied to an ultra-light aircraft, in order to improve aerodynamic efficiency and flight control. Given the size (less than a meter of wingspan) and the low weight of this aircraft, SMAs were more appropriate than traditional actuators, thanks to their high strength, low weight, and reduced footprint. The structure was made of balsa wood with nylon sticks; wings were composed of two parts, a rigid front element, and a deformable rear. SMA actuators were used in pairs, and so each wire acted in opposition to another: this concept is often found in the literature and is called the antagonistic configuration. Benavides and Correa (2004) realized a wing section based on the symmetric Gottingen 776 airfoil, with a deformable trailing edge actuated by SMA wires. Wind tunnel tests verified the improvement in terms of lift-to-drag ratio and lift coefficient of this solution with respect to a traditional one. The wing section used six

NiTiNOL wires that could pull, upon electrical activation, the upper part of the wing trailing edge downward. In a different application, Madden et al. (2004) varied the camber of a marine propeller using polypyrrole actuators. These actuators behave like artificial muscle by extending and contracting with voltage, and promise to provide the actuation forces and strains required to change the camber of the propeller foils. Two polypyrrole actuators operated as an antagonistic pair to operate tendons attached to the trailing edge of the foil. In order to achieve the desired degree of strain, two concepts were compared; the two actuators were arranged in the form of a bimetallic strip so as to complement each other, or a lever arrangement was used. Both these methods were examined by experimental testing. Yang et al. (2006) presented the camber change of an airfoil via the use of SMAs. A symmetric wing containing four pairs of SMA actuators was used in the wind tunnel experiments. The smooth change in the camber caused a significant lift increase with a small drag penalty. Mirone (2007) and Mirone and Pellegrino (2009) investigated the ability of the wing to modify its crosssection (assuming the shape of two different airfoils) and the possibility of deflecting the profiles near the trailing edge in order to obtain hingeless control surfaces. Oneway SMA wires coupled to springs provided the actuation. The points to be actuated along the profiles and the displacements to be imposed were selected so that they satisfactorily approximated the change from one airfoil to the other and resulted in an adequate deflection of the control surface. The actuators and their performance were designed to guarantee adequate wing stiffness in order to prevent excessive deformations and undesired airfoil shape variations due to aerodynamic loads. Two prototypes were realized, incorporating the variable airfoil and the hingeless aileron features respectively, and the verification of their shapes in both the actuated and non-actuated states. The comparison of the real wing profiles with the theoretical airfoil coordinates showed a good performance of the actuation system. A significant trailing edge displacement, close to 10% of the wing chord, was achieved in the experimental tests. Song and Ma (2007) used SMA wires to control the flap movement of a model airplane wing mainly to achieve weight reduction. Two SMA actuators were used in an antagonistic way. A sliding mode-based non-linear robust controller was designed and implemented on a realtime data acquisition and control platform to control the position of the flap. Experimental results show that high control accuracy has been achieved for both the position regulation and tracking tasks, even with uncertainties and disturbances. Brailovski et al. (2008) considered a morphing wing concept designed for subsonic cruise flight conditions combining three principal sub-systems: a flexible skin,

862

S. BARBARINO ET AL.

a rigid interior structure, and an actuator group located inside the wingbox. Using a weighted evaluation of the aerodynamic and mechanical performances in the framework of a multi-criteria optimization procedure, an adaptive structure consisting of a four-ply flexible skin powered by two individually controlled actuators was designed. The forces these actuators must provide in order to fit target wing profiles are calculated for a given flexible structure and given flight conditions. Each actuator includes linear SMA active elements with one extremity connected, through a cam transmission system, to the flexible skins and other extremity connected to the rigid structure of the wing through a bias spring. To meet the functional requirements of the application, the geometry (length and cross-section) of the SMA active elements and the bias spring characteristics are calculated. Peel et al. (2008, 2009) constructed a variable-camber morphing wing using elastomeric composites as the internal actuators. The skin of the wing was a carbon fiber and polyurethane elastomeric composite that was able to flex in the chordwise direction. The actuators were RMAs. When pressurized with air, the RMAs contract with a force generally proportional to the applied pressure. Similar to hydraulic systems, these actuators require pneumatic systems, which typically have a weight penalty of the supporting hardware (compressor and accumulator, or pressure storage vessel, solenoids, control circuitry). The RMA actuators were connected to the leading and trailing edges of the wing; hence the contraction would cause an increase in camber. Two generations of prototypes were described with significant improvements made on the second generation. A very simple morphing wing was fabricated in phase one: the nose was able to elastically camber down by about 25 and the tail by 20 . The second-generation wing fabrication methodology showed smooth elastic cambering and no buckling or waviness in the skins. In this case, the LE and the TE achieved 23 and 15 deflections, respectively. Chou and Philen (2008) presented morphing skins employing pressurized flexible matrix composites (FMCs) for trailing-edge flap control of a glider wing. The FMC skins consisted of embedded pressurized composite tubes in an elastomeric resin. A mathematical model for a single pressurized FMC composite tube was developed and the rule of mixtures was employed to find the effective properties of the multi-tube active skin. A 2D vortex panel method estimated the aerodynamic loading on the trailing-edge flaps. Popov et al. (2008) presented an adaptive wing concept with an upper surface that deflected under single-point control. The flexible skin was placed at the laminarturbulent boundary layer transition point, and different functions were tested to find whether the transition could be shifted toward the trailing edge and

whether drag could be reduced. In addition to static deflections, sinusoidal deflections are evaluated. Seow et al. (2008) presented a study on morphing flaps actuated by SMA wires, including a theoretical analysis of the aerodynamic loading. A wing prototype was designed and fabricated, and the actuation of the flap was demonstrated, achieving an actuation angle of 5.2 . A second prototype achieved 15 deflection. Flexible skins were also explored and experimented. Finally, the controllability of the prototype was investigated, including the measurement and estimation of power consumption, and the dependence of heating and cooling time on input variations. Abdullah et al. (2009) used SMAs to alter the shape of an airfoil to increase the L/D ratio throughout the flight regime of a UAV. A flexible skin was also utilized for variablecamber control. The aerodynamic effect of camber location and magnitude was analyzed using a 2D panelmethod. The researchers found that shifting the location of maximum camber thickness from 10% to 50% of the chord, and increasing the camber from 1% to 5%, maximized L/D for low angles of attack. Ideally, the wing would use this range to accommodate multiple flight conditions. The objective was predominantly cruise efficiency rather than control, and hence the response time of the SMA was considered to be acceptable. Barbarino et al. (2009a) investigated an active wing (via SMAs) that deformed its shape to provide a bumped wing geometry. This variation in the wing thickness is advantageous for transonic flight, but also has benefits in subsonic flight. Benchtop experiments validated the efficacy of the concept. Barbarino et al. (2007, 2008, 2009b) introduced several morphing trailing edge architectures to replace a conventional flap device. A compliant rib structure was designed, based on SMA actuators exhibiting structural potential (i.e., bearing the external aerodynamic loads). Numerical results, achieved through an FE approach, were presented in terms of trailing edge-induced displacement and morphed shape. Barbarino et al. (2010a) proposed a flap architecture for a variable-camber trailing edge, whose reference geometry was based on a full-scale wing for a regional transport aircraft. The compliant rib was based on a truss-like structure where some members were active rods made of SMA. The layout of the structure was obtained using a preliminary optimization process, incorporating practical constraints, by focusing on a basic truss-like element and its repetition and position within the overall truss. The structural performance was estimated using FE analysis. The SMA behavior was modeled using a dedicated routine to evaluate the internal stress state and the minimum activation temperature. The design fulfills a number of key requirements: a smooth actuation along the chord; morphed shapes that assure the optimal aerodynamic load distribution for high lift; light weight; and low mechanical

Morphing Aircraft

863

complexity. The design features of the architecture were investigated and the requirements of the morphing skin discussed. Icardi and Ferraro (2009) designed morphing wings with SMA torsion tubes and independent SMA wires for wing camber control. The analysis delivered stressstrain curves for various types of SMA wires and stresses on structural wing components of the indicated design, including a supporting wingbox section and surface material. Their theoretical model predicted that for a 1000 kg UAV with 1.2 m (mean) chord and 3 m span, the camber varied from 5 to 15.5 with SMA torsion tubes delivering 150200 Nm and drawing 1.2 kW of power. Lv et al. (2009) presented a brief survey of adaptive materials and structures research in China. The growing application of SMP, SMA, electro-active polymer, MFC, and piezoelectric materials for aerospace were discussed. Functional composite filler materials such as conductive nanoparticles, short carbon fiber, and ferroelectric ceramics were introduced. The authors provided an example of an SMA torsion actuator based on NiTi wires and thin-walled tubes used for a morphing wing. The angle of attack of a wing model with the SMA torsion actuator could be changed continuously up to 15 in 1 s. In addition, a novel smart wing rib was presented, capable of reaching 10 deflection. Piezoelectric Actuation for Camber Piezoelectric materials offer relatively high force output in a wide frequency bandwidth. Although the strain output is relatively low (compared to SMAs), the fast response of PZTs (initially) caused interest in terms of vibration control. Many researchers focused on the application of piezoelectrics to rotor systems to improve their performance and effectiveness. Several smart rotor concepts have been developed: leadingand trailing-edge flaps actuated with smart material actuators, controllable camber/twist blades with embedded piezoelectric elements/fibers, and active blade tips actuated with tailored smart actuators. For flap actuation, actuators range from piezobimorphs (Spangler and Hall, 1990; Koratkar and Chopra, 2000, 2001), piezostacks (Janker et al., 1999; Lee, 1999; Lee and Chopra, 2001a, b; Straub et al. 2001, 2009), and piezoelectric/ magnetostrictive-induced composite-coupled systems (Derham and Hagood, 1996; Rogers and Hagood, 1997; Bernhard and Chopra, 2001; Cesnik and Shin, 2001; Cesnik et al., 2001; Shin, 2001; Shin et al., 2002; Kovalovs et al., 2007). Straub et al. (2001, 2009) presented Boeings smart active rotor, with piezoelectricactuated blade flaps, which was tested in a wind tunnel at NASA Ames Research Center in 2008, demonstrating over 80% reduction in vibration and the potential for zero vibration at the hub.

Static control of aerodynamic surfaces using piezoelectrics started in early 1990s. Lazarus et al. (1991) examined the feasibility of using representative box wing adaptive structures for static aeroelastic control. Greater control authority and a lower weight penalty are achieved using adaptive aeroelastic structures for a variety of wing designs. In addition, the important parameters associated with inducing curvature and twist, to alter the lift on the airfoil, were determined: the airfoil thickness ratio, the actuation strain produced by the induced strain actuators, and the relative stiffness ratio of the actuator to the wing skin for both camber and twist control. Pinkerton and Moses (1997) discussed the feasibility of controlling the wing geometry employing a piezoelectric actuator known as thin-layer composite-unimorph ferroelectric driver and sensor (known as the THUNDER actuator). This morphing skin was used to create a bubble in the skin of the airfoil near the leading edge to extend the region of attached flow further toward the trailing edge. Such a capability would expand the airspeeds over which the aircraft could maintain flight in general, but more importantly efficient flight. Hysteresis non-linearity was observed in the voltage-to-displacement relationship. Geissler et al. (2000) adopted piezoelectric materials as the actuation element for a morphing leading edge; however, in this case, the leading edge is an independent element able to rotate around an internal hinge. This technique was conceived and adopted for rotorcraft morphing, due to the small size of actuators. Munday and Jacob (2001) developed a wing with conformal curvature, characterized by a piezoelectric actuator internally mounted in a position to alter the upper surface shape of the airfoil, resulting in a variation of the effective curvature. The concept allowed for the modification of airfoil thickness and position of its maximum value. The piezoelectric actuator (THUNDER) allows for a continuously variable movement, thus resulting in both static and dynamic shape change controls. In 2001, Barrett (2002) introduced a convertible coleopter UAV, capable of hovering like a helicopter, and then transition nearly 90 to fly like an airplane. The XQ-138 aircraft achieved yaw control through deflections of its turning vane flaps, and pitch and roll control were maintained by a set of cascading grid fins. The aircraft was later fitted with the post-buckled-precompression (PBP) actuators (Barrett and Tiso, 2004). Wang et al. (2001) presented results from the Smart Wing Phase 2 program which demonstrated high-rate actuation of hingeless control surfaces using several smart material-based actuators. In the first study, several actuation concepts with different transducers were modeled and analyzed. These concepts included distributed piezoelectric stack actuators with and without hydraulic amplifiers and pumps, antagonistic tendon

864

S. BARBARINO ET AL.

actuation, and eccentuation. The transducers selected for the trade studies included piezoelectric ultrasonic motors, actively cooled SMA, ferromagnetic SMA, and stacks made from piezoelectric ceramic wafer, piezoelectric single crystal wafer, irradiated PVDF-TrFE film, and dielectric elastomer film. The analyses showed that distributed polymer stacks provided the most elegant solution, but eccentuation was deemed the most realistic and lowest risk approach to attaining the program goals. A common issue to all the concepts was the structural stiffness that the actuators worked against. This was resolved in the second study by developing a flexcore elastomeric skin trailing edge structure with eccentuation using high-power ultrasonic motors. Grohmann et al. (2006, 2008) presented the active trailing edge and active twist concepts applied to helicopter rotor blades. The sizing and placement of piezocomposite patches were optimized to obtain the desired wing weight and pitching moment. A piezo-composite actuator was developed with a relatively high strain and force output. The recent interest in smaller and lighter aircraft has driven the use of smart materials for flight and flow control. For example, field-deployable aircraft have flexible wings that can be folded during transportation, and they can be unfolded for operation. These compliant wings can be realized with the integration of smart materials. For smart material-actuated devices, operating a relatively compliant, thin structure (desirable for piezoelectric actuators) in situations where there are relatively high external forces is challenging. Establishing a wing configuration that is stiff enough to prevent flutter and divergence, but compliant enough to allow the range of available motion is the central challenge in developing a piezocomposite airfoil. Novel methods of supporting the actuator can take advantage of aerodynamic loads to reduce control input moments and increase control effectiveness. In the following paragraphs, small unmanned (and/or remotely piloted) fixed-wing, rotary, and ducted fan aircraft are considered that employ piezoelectric materials for aerodynamic control. The 2002 Virginia Tech Morphing Wing Design Team (Eggleston et al., 2002) experimented with the use of piezoelectric ceramics (THUNDER), shape memory alloys (NiTi), and conventional servomotors. A flying wing planform was chosen as the testbed for this project. A morphing specific VLM code (MorphVLM) was developed to obtain aerodynamic loads. Theoretical analysis was conducted on various trailing-edge shapes in order to optimize values for a trimmed CLmax. Wind tunnel testing analyzed the aerodynamic characteristics of the morphing and conventional models. The project concluded that morphing technologies have great potential to improve the aerodynamics of aircraft. Barrett et al. (2005) employed piezoelectric elements along with elastic elements to magnify control

deflections and forces. The so-called PBP concept was employed as guide vanes in a small rotary aircraft. The PBP concept, in its earliest incarnation, was primarily intended to increase the coupling coefficient exhibited by piezoelectric transducer elements (Lesieutre and Davis, 1997). Early experimentation showed that apparent coupling coefficients approaching one could be achieved by axially loading bending elements with forces that approached the buckling load of the beam. Vos et al. (2007a, b, 2008a) improved the PBP concept for aerodynamic applications. An elastic skin covering the outside of the wing generated the axial precompression in the piezoelectric elements and also served as an aerodynamic surface. Trailing-edge deflections of around 3.1 could be attained up to 34 Hz. Roll control authority was increased on a 1.4-m span unmanned air vehicle. The experiments showed an increase in roll authority while reducing weight, part-count, and power consumption. Bilgen et al. (2007, 2009) presented a new application for piezo-composite actuators on a 0.76-m wingspan morphing wing air vehicle with approximately 0.815 kg total weight. In this application, two MFC patches were bonded to the wings of a small demonstration vehicle to change the camber of the wing. The change in the wing camber provided adequate roll control authority in the wind tunnel and also in flight. The aircraft demonstrated that lightweight, conformal actuators can be used as primary control surfaces on an aircraft. All electronics, including the MFC power electronics, were powered by an 11.1 V Lithium polymer battery, which is a common choice for remotely controlled aircraft. The aircraft used a 150-W brushless motor for thrust; in contrast, the MFC power electronics consumed only 3 W during peak actuation. In a different application, Bilgen et al. (2009) presented a variable-camber airfoil design (employing MFCs) intended for a ducted fan aircraft. The study focused on 2D aerodynamic and structural response characterization under aerodynamic loads for circular-arc airfoils with variable pinned boundary conditions. A parametric study of fluidstructure interaction was employed to find pin locations along the chord-wise direction that resulted in high lift generation. Wind tunnel experiments were conducted on a 1.0% thick MFC-actuated bimorph airfoil that was simply supported at 5% and 50% chord. Aerodynamic and structural performance results were given and nonlinear effects due to aerodynamic and piezoceramic hysteresis were identified and discussed. A lift coefficient change of 1.46 was observed purely due to voltage actuation. A maximum 2D L/D ratio of 17.8 was recorded. Paradies and Ciresa (2009) implemented MFCs as actuators in an active composite wing. The research designed a wing for a UAV with a thin profile and integrated the piezoelectric elements for roll control. The design and its optimization were based on a fully

Morphing Aircraft

865

coupled structural fluid dynamics model that implemented constraints from available materials and manufacturing. A scaled prototype wing was manufactured. The design model was validated with static and preliminary dynamic tests of the prototype wing. The qualitative agreement between the numerical model and experiments was good. Dynamic tests were also performed on a sandwich wing of the same size with conventional aileron control for comparison. Even though the roll moment generated by the active wing was lower, it proved sufficient for the intended control of the UAV. The active wing with piezoelectric flight control was one of the first examples where such a design had been optimized and the numerical model had been validated in experiments. Wickramasinghe et al. (2009) presented the design and verification of a smart wing for a UAV. The proposed smart wing structure consisted of a composite spar and ailerons that had bimorph active ribs consisting of MFC actuators. The actuation was enhanced by preloading the piezoceramic fiber actuators with a compressive axial load. The preload was exerted on the actuators through a passive latex or electro active polymer skin that wrapped around the airfoil. Bilgen et al. (2010a, b) presented research to enable solid-state aerodynamic force generation in highdynamic pressure airflow. A bi-directional variablecamber airfoil employing MFCs was presented. The airfoil employed two active surfaces and a single solidstate four-bar (box) mechanism as the internal structure. The unique choice of boundary conditions allowed variable and smooth deformation in both directions from a flat camber line. The paper focused on actuation modeling and response characterization under aerodynamic loads. First, a parametric study of the aerodynamic response was used to optimize the kinematic parameters of the airfoil. The concept was fabricated with MFC actuators in a bimorph configuration for the active surfaces. Wind tunnel experiments were conducted on a 12.6% maximum thickness airfoil. Non-linear effects due to aerodynamic and piezoceramic hysteresis were identified and discussed. A lift coefficient change of 1.54 was observed and the results compared to conventional, zero-camber NACA and other airfoils. The variable-camber airfoil produced 72% higher lift curve slope when compared to the lift curve slope of a symmetric NACA 0009 airfoil. Bilgen et al. (2011a) presented a surface-actuated variable-camber morphing airfoil employing MFC actuators. The continuity of the airfoil surface was achieved using a single substrate that wraps around the airfoil shape. This substrate forms the surface of the airfoil and serves as the host material for the two cascading bimorph actuators. A parametric study of the fluidstructure interaction problem optimized the geometric parameters and the boundary conditions of the variable-camber airfoil. The coupled treatment of the fluidstructure interaction allowed a design that

can sustain large aerodynamic loads. The paper identified the effects of four important structural parameters to achieve the highest possible lift coefficient and lift-todrag ratio. The 2010 Virginia Tech Wing Morphing Design Team (Butt et al., 2010a, b; Bilgen et al., 2011b) developed a completely servo-less, operational, wind-tunnel, and flight-tested remotely piloted aircraft. The team members were able to develop lightweight control surfaces and the necessary driving high-voltage DCDC converters, culminating in a landmark first flight on April 29, 2010. The morphing design replaced all the traditional servomotor-controlled surfaces with MFC-actuated surfaces in an effective manner and all systems were powered with a single Lithium polymer battery. This vehicle became the first fully solid-state piezoelectric controlled, non-tethered, flight-tested aircraft. The MFC-actuated control surfaces did not employ any rotating, moving, or multi-piece parts or mechanisms. Kinematic or compliant mechanical amplification mechanisms were not employed, which resulted in a virtually solid-state and practical aerodynamic control surface. Summary It is clear that the airfoil morphing, and more specifically camber morphing, is the dominant research topic in subsonic aerodynamic applications when compared to the planform and out-of-plane morphing methods. The research in airfoil morphing (which includes camber and thickness distribution changes) is dominated by the camber morphing concepts. The categorization of camber morphing by the method of actuation causes a relatively distinct division of the size of the aircraft where (1) conventional actuators are used in all aircraft sizes (except sub-MAV scale), (2) SMAs are used in rotorcraft blades and UAVs, and (3) PZTs are used mostly in small UAVs and MAVs. The most common actuation method in airfoil morphing is conventional (lumped) actuators. Airfoil morphing through conventional actuation has one example in 1973 with the rest of the examples starting from 1984. Of the total number of academic works, less than a third of these papers provide wind tunnel and flight tests. The applications are broad, ranging from transport and fighter aircraft to small UAVs and MAVs. The lumped actuation methods are servo- and ultrasonic motors, and pneumatic and hydraulic devices. Observing the number of papers, the range of actuators and applications, it is important to note that these conventional (and lumped) actuation methods appear to be able to provide the force and frequency requirements for subsonic aerodynamic flight control. It is also observed that the conventional actuators are almost solely used in morphing of fixed-wing aircraft (and rarely in airfoil morphing of blades of rotary-wing aircraft).

866

S. BARBARINO ET AL.

The second most popular actuation method is SMAlike actuation where relatively large strains can be applied by the contracting active elements. Although SMAs are smart material actuators, due to their slow frequency response (inherent due to the thermodynamic heating and cooling behavior), they are used with conventional logistical systems. In most applications, where a low-temperature thermal sink is not available, SMAs are typically employed with relatively complex mechanical components (such as active fluidic cooling systems) to speed up their thermodynamic behavior. Applications of SMAs (and similar devices) vary between fighter aircraft, transport aircraft, and UAVs, both for fixed-wing and rotary-wing configurations. The research has the first example in 1994, and other examples of wind tunnel testing in the 19942006 period. Only one flight test has been achieved, where the application was a rotary-wing UAV. Almost all the research employs theoretical analysis, whereas only half employs experimental (benchtop) validation. The research has been directed mostly toward the fixed-wing application since 2003. The third actuation method for inducing camber change is via PZTs (which has typical characteristics of high frequency, high load, and low strain). The research in this area has examples since 1990, almost all having theoretical analysis and benchtop and wind tunnel validation. There are four flight-tested aircrafts which are all small UAVs. Equal attention is given to rotary- and fixed-wing aircrafts. PZT actuation is directed mostly to overall camber change of the airfoil, and specifically the camber of the trailing section. No MAV applications are presented for PZT actuation (mainly because flappingwing aircrafts are not considered in this review). PZTs are good candidates for actuation in small UAVs and MAVs due to the required high-frequency response. Overall, the lack of benchtop and wind tunnel validation of the proposed airfoil morphing concepts is consistent for all actuation methods in the topic of camber morphing (for subsonic aerodynamic applications). Due to their inherent frequency response, SMAs are more popular (and probably more attractive in the future) for large vehicles such as transport aircrafts, where PZTs are more popular (and desired) for small UAVs and MAVs. For all applications and actuation methods, the importance of the skin is usually overlooked because the proposed concepts usually deal with 2D aerodynamic configurations. This could be one reason why there are a significant number of wind tunnel evaluations, however only a few flight tests. CONCLUSIONS The field of shape morphing aircraft has attracted the attention of hundreds of research groups worldwide. Although many interesting concepts have been synthesized, few have progressed to wing tunnel testing, and

even fewer have ever flown. Wing morphing is a promising technology, because it allows the aerodynamic potential of an aircraft wing to be explored, by adapting the wing shape for several flight conditions encountered in a typical mission profile. Moreover, the aero-elastic deformations can increase the performance and maneuverability, and improve the structural efficiency. Variable sweep has proven successful, particularly to enable military aircraft to fly at supersonic speeds, albeit with a large weight penalty. The current trend for highly efficient and green aircraft makes this compromise less likely to be acceptable, calling for innovative morphing designs able to provide more benefits and fewer drawbacks. For military applications, the current level of performance required by the next-generation vehicles cannot sustain this trade-off. In general, any successful wing morphing system must overcome the weight penalty due to the additional actuation systems. Compared to supersonic aircrafts, small or low-speed vehicles require more dramatic wing variations to produce a noticeable and practical change in their aerodynamic properties. The explosive growth of satellite services during the past few years has made UAVs the technology of choice for many routine applications such as border patrol, environmental monitoring, meteorology, military operations, and search and rescue. For this and other reasons (such as lower production costs, lower safety and certification requirements, and lower aerodynamic loads), many investigations on wing morphing are focused toward small or radio-controlled aircraft, i.e., UAVs. This also offers a great opportunity to showcase and test successful designs at an early stage, and to attract industry attention to develop new technologies for largescale vehicles. However, manufacturers are still too skeptical of the benefits to adopt morphing technologies in the near future, as many developed concepts have a technology readiness level that is still too low. Recent advances in smart materials research, including developments in actuation technology, constitutive laws and modeling, optimization and control, and failure prediction, demand more purposeful steps to progress variable-geometry small aircraft. The blending of morphing and smart structures (defined as structures capable of sensing the external environment, processing the information and reacting accordingly) seems mandatory to enable innovative solutions to be realizable and competitive with traditional designs, and overcome the penalties associated with current morphing applications. This integrated approach requires multi-disciplinary thinking from the early stages of development, which significantly increases the overall complexity even of the preliminary design. Design interactions are more problematic with morphing vehicles. The future of morphing is uncertain. Technology programs (beyond basic research) need to be matched to

Morphing Aircraft

867

capability gaps to insure relevance and funding. Morphing does show promise for several types of missions, but to say there is a compelling case for morphing is perhaps an overstatement. Morphing should be viewed as a design option to be incorporated in a specific vehicle if justified by system-level benefits achieved for the costs incurred. Morphing as a suite of technologies is not flight ready. Much work is needed in maturing component technologies such as skins (flexible or stretchable while carrying loads), actuators/mechanisms (distributed and capable of supporting part of the external loads), and control theory (primary flight and actuation) for morphing to be truly realized. But for all these efforts to pay off, morphing technology needs a transition program for the aerospace community to take the technology seriously and include it as a design option. While morphing on vehicles exceeding several thousand pounds gross weight may not be practical in the near term due to the low technology readiness levels, opportunities exists on smaller unmanned aircraft or missiles where current or near-term technology can be applied to achieve morphing. According to studies at NASA, it will take another 2030 years before skies could be filled by aircraft more similar to birds, having wings without discrete control surfaces that can change their shape in a smart way. ACKNOWLEDGMENTS The authors acknowledge funding from the European Research Council through grant number 247045 entitled Optimisation of Multiscale Structures with Applications to Morphing Aircraft. REFERENCES
Abdullah, E.J., Bill, C. and Watkins, S. 2009. Application of Smart Materials for Adaptive Airfoil Control, In: Proceedings of 47th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, 58 January, Orlando, FL, AIAA 2009-1359 (11 pp). Abdulrahim, M. and Lind, R. 2004. Flight Testing and Response Characteristics of a Variable Gull-Wing Morphing Aircraft, In: Proceedings of AIAA Guidance, Navigation, and Control Conference and Exhibit, 1619 August, Providence, RI, AIAA, 2004-5113, pp. 16641679. Abdulrahim, M. and Lind, R. 2006. Using Avian Morphology to Enhance Aircraft Maneuverability, In: Proceedings of AIAA Atmospheric Flight Mechanics Conference and Exhibit, 2124 August, Keystone, CO, AIAA 2006-6643. AeroVisions Inc. 2011. Available at: http://www.canosoarus.com/ 05UMAAV/UMAAV01.htm (accessed date May 22, 2011). AFRL Vehicle Research Section, Air Launched Integrated Countermeasure, Expendable (ALICE). 2011. Available at http://www.nrl.navy.mil/vrs/factsheets/index.php (accessed date May 22, 2011). Ajaj, R.M., Friswell, M.I., Dettmer, W.G., Allegri, G. and Isikveren, A.T. 2011a. Conceptual Modelling on Adaptive Torsion Wing Structure, 19th AIAA/ASME/AHS Adaptive Structures Conference, Denver, Colorado, AIAA 20111883 (15 pp).

Ajaj, R.M., Friswell, M.I., Smith D.D. and Isikveren, A.T. 2011b. Roll Control of a UAV Using an Adaptive Torsion Structure 7th AIAA Multidisciplinary Design Optimization Specialist Conference, Denver, Colorado, AIAA 20111884 (15 pp). Alasty, A., Alemohammad, S.H., Khiabani, R.H. and Khalighi, Y. 2004. Maneuverability Improvement for an Ultra Light Airplane Model Using Variable Shape Wing, In: Proceedings of AIAA Atmospheric Flight Mechanics Conference and Exhibit, 1619 August, Providence, RI, AIAA 2004-4831 (10 pp). Albertani, R., Hubner, J.P., Ifju, P.G., Lind, R. and Jackowski, J. 2004. Experimental Aerodynamics of Micro Air Vehicles, In: SAE World Aviation Congress and Exhibition, 3 November, Reno, NVPaper 04AER-8. Alemayehu, D., Leng, M., McNulty, R., Mulloy, M., Somero, R. and Tenorio, C., 2004. Virginia Tech Morphing Wing Team Fall 2004 Final Report. Aerospace and Mechanical Engineering Departments, Virginia Tech, Blacksburg, VA. Available at: http://www.aoe.vt.edu/~mason/Mason_f/MorphF04Rpt.pdf (accessed date May 22, 2011). Alemayehu, D., Leng, M., McNulty, R., Mulloy, M., Somero, R. and Tenorio, C., 2005. Virginia Tech Morphing Wing Team Spring 2005 Final Report. Aerospace and Mechanical Engineering Departments, Virginia Tech, Blacksburg, VA. Available at: http://www.aoe.vt.edu/~mason/Mason_f/MorphS05Rpt.pdf (accessed date May 22, 2011). Amprikidis, M. and Cooper, J.E. 2003. Development of Smart Spars for Active Aeroelastic Structures, In: Proceedings of 44th AIAA/ ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, 710 April, Norfolk, VA, AIAA 20031799. Amprikidis, M., Cooper, J.E. and Sensburg, O. 2004. Development of An Adaptive Stiffness All-Moving Vertical Tail, In: Proceedings of 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 1922 April, Palm Spring, CA, AIAA 2004-1883 (13 pp). Anderson, J.D. 2000. Introduction to Flight, 4 edn, McGraw Hill Book Company, New York. Andersen, G.R., Cowan, D.L. and Piatak, D.J. 2007. Aeroelastic Modeling, Analysis and Testing of a Morphing Wing Structure, In: 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-1734 (15 pp). Arrison, L., Birocco, K., Gaylord, C., Herndon, B., Manion, K. and Metheny, M. 2003. 2002-2003 AE/ME Morphing Wing Design, Spring Semester Final Report, Virginia Tech Aerospace Engineering Senior Design Project. Austin, F., Siclari, M.J., Van Nostrand, W., Weisensel, G.N., Kottamasu, V. and Volpe, G. 1997. Comparison of Smart Wing Concepts for Transonic Cruise Drag Reduction, In Proceedings of SPIE 1997 Smart Structures and Materials 1997: Industrial and Commercial Applications of Smart Structures Technologies, 4 March San Diego, CA, 3044:3340. Bae, J.-S., Seigler, T.M. and Inman, D.J. 2005. Aerodynamic and Static Aeroelastic Characteristics of a Variable-Span Morphing Wing, Journal of Aircraft, 42:528534. Bailly, J. and Delrieux, Y. 2009. Improvement of Noise Reduction and Performance for a Helicopter Model Rotor Blade by Active Twist Actuation, In: Proceedings of 35th European Rotorcraft Forum, 2225 September, Hamburg, Germany, Article 1157. Baker, D. and Friswell, M.I. 2006. Design of a Compliant Aerofoil using Topology Optimisation, In: Proceedings of International Workshop on Smart Materials and Structures (Cansmart 2006), 1213 October, Toronto, Canada (9 pp). Baker, D. and Friswell, M.I. 2009. Determinate Structures for Wing Camber Control, Smart Materials and Structures, 18:035014. Baker, D., Friswell, M.I. and Lieven, N.A.J. 2005. Active Truss Structures for Wing Morphing, In: Proceedings of II ECCOMAS Thematic Conference on Smart Structures and Materials, 1821 July, Lisbon, Portugal (11 pp).

868

S. BARBARINO ET AL.
Barrett, R.M. and Stutts, J.C. 1997. Design and Testing of a 1/12th Scale Solid State Adaptive Rotor, Journal of Smart Materials and Structures, 6:491497. Barrett, R.M. and Stutts, J.C. 1998. Development of a Piezoceramic Flight Control Surface Actuator for Highly Compressed Munitions, In: Proceedings of 39th Structures, Structural Dynamics and Materials Conference, 2023 April, Long Beach, CA, AIAA 98-2034, pp. 28072813. Barrett, R.M. and Tiso, P. 2004. PBP Adaptive Actuators and Embodiments, International Patent Filing from the Faculty of Aerospace Engineering, Technical University of Delft, Netherland. Barrett, R.M., Vos, R., Tiso, P. and De Breuker, R. 2005. PostBuckled Precompressed (PBP) Actuators: Enhancing VTOL Autonomous High Speed MAVs, In: Proceedings of 46th AIAA/ASME/ASCE/AHS/ASC Structure, Structural Dynamics & Materials Conference, 1821 April, Austin, TX, AIAA 20052113 (13 pp). Bartley-Cho, J.D., Wang, D.P., Martin, C.A., Kudva, J.N. and West, M.N. 2004. Development of High-Rate, Adaptive Trailing Edge Control Surface for the Smart Wing Phase 2 Wind Tunnel Model, Journal of Intelligent Material Systems and Structures, 15:279291. Beauchamp, C.H. and Nedderman Jr., W.H. 2001. Patent: Controllable Camber Windmill Blades. Office of Naval Research. Benavides, J.C. and Correa, G. 2004. Morphing Wing Design Using Nitinol Wire. University of Missouri-Rolla (UMR), Intelligent System Center, Rolla, MO. Available at: http://isc.mst.edu/reu/ 2004indprojects/2004-6.html (accessed date May 22, 2011). Bernhard, A.P.F. and Chopra, I. 2001. Analysis of a BendingTorsion Coupled Actuator for a Smart Rotor with Active Blade Tips, Smart Materials and Structures, 10:3552. Bharti, S., Frecker, M.I., Lesieutre, G. and Browne, J. 2007. Tendon Actuated cellular Mechanisms for Morphing Aircraft Wing, In: Proceedings of SPIE Modeling, Signal Processing, and Control for Smart Structures, 19 March, San Diego, CA, Vol. 6523, doi:10.1117/12.715855. Bilgen, O., Butt, L.M., Day, S.R., Sossi, C.A., Weaver, J.P., Wolek, A., Mason, W.H. and Inman, D.J. 2011b. A Novel Unmanned Aircraft with Solid-State Control Surfaces: Analysis and Flight Demonstration, In: Proceedings of 52nd AIAA/ASME/ASCE/ AHS/ASC Structures, Structural Dynamics, and Materials, 47 April, Denver, CO, AIAA 2011-2071 (24 pp). Bilgen, O., Friswell, M.I., Kochersberger, K.B. and Inman, D.J. 2011a. Surface Actuated Variable-Camber and Variable-Twist Morphing Wings Using Piezocomposites, In: Proceedings of 52nd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials, 47 April, Denver, CO, AIAA 20112072. Bilgen, O., Kochersberger, K.B., Diggs, E.C., Kurdila, A.J. and Inman, D.J. 2007. Morphing Wing Micro-Air-Vehicles Via Macro-Fiber-Composite Actuators, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-1785 (17 pp). Bilgen, O., Kochersberger, K.B. and Inman, D.J. 2009. Macro-Fiber Composite Actuators for a Swept Wing Unmanned Aircraft, Aeronautical Journal, Special issue on Flight Structures Fundamental Research in the USA, 113:385395. Bilgen, O., Kochersberger, K.B. and Inman, D.J. 2010a. MacroFiber Composite Actuated Simply-Supported Thin Morphing Airfoils, Smart Material and Structures, 19:055010 (11pp). Bilgen, O., Kochersberger, K.B. and Inman, D.J. 2010b. Novel, BiDirectional, Variable Camber Airfoil via Macro-Fiber Composite Actuators, AIAA Journal of Aircraft, 47:303314. Blondeau, J. and Pines, D. 2007. Design and Testing of a Pneumatic Telescopic wing for Unmanned Aerial Vehicles, Journal of Aircraft, 44:10881099. Blondeau, J., Richeson, J. and Pines, D.J. 2003. Design, Development and Testing of a Morphing Aspect Ratio Wing

Barbarino, S., Ameduri, S., Lecce, L. and Concilio, A. 2009a. Wing Shape Control through an SMA-Based Device, Journal of Intelligent Material Systems and Structures, 20:283296. Barbarino, S., Ameduri, S. and Pecora, R. 2007. Wing Chamber Control Architectures based on SMA: Numerical Investigations, In: Du, S., Leng, J. and Asundi, A.K. (eds), Proceedings of SPIE International Conference on Smart Materials and Nanotechnology in Engineering (SMN2007), 14 July, Harbin, China, Vol. 6423, p. 64231E. Barbarino, S., Concilio, A., Pecora, R., Ameduri, S. and Lecce, L. 2008. Design of an Actuation Architecture based on SMA Technology for Wing Shape Control, In: Proceedings of Actuator 2008 Conference, 911 June, Bremen, Germany, p. 144. Barbarino, S., Dettmer, W.G. and Friswell, M.I. 2010a. Morphing Trailing Edges with Shape Memory Alloy Rods, In: Proceedings of 21st International Conference on Adaptive Structures and Technologies (ICAST), 46 October, State College, PA (17 pp). Barbarino, S., Gandhi, F. and Webster, S.D. 2010b. Design of Extendable Chord Sections for Morphing Helicopter Rotor Blades, In: Proceedings of ASME 2010 Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS2010), 28 September1 October, Philadelphia, PASMASIS2010-3668 (14 pp). Barbarino, S., Pecora, R., Lecce, L., Concilio, A., Ameduri, S. and Calvi, E. 2009b. A Novel SMA-Based Concept for Airfoil Structural Morphing, Journal of Materials Engineering and Performance, 18:696705. Barrett, R.M. 1990. Intelligent Rotor Blade Actuation through Directionally Attached Piezoelectric Crystals, In: Proceedings of 46th American Helicopter Society National Conference and Forum, 2123 May, Washington, DC. Barrett, R.M. 1992a. Actuation Strain Decoupling trough Enhanced Directional Attachment in Plates and Aerodynamic Surfaces, In: Proceedings of First European Conference on Smart Structures and Material, 1214 May, Glasgow, Scotland. Barrett, R.M. 1992b. High Bandwidth Electric Rotor Blade Actuator Study, Phase 1 SBIR Proposal, The US Army Aviation Systems Command, St. Louis, MO. Barrett, R.M. 1996. Active Aeroelastic Tailoring of an Adaptive Flexspar Stabilator, Smart Materials and Structures, 5:723730. Barrett, R.M. 2002. Convertible Vertical Take-Off and Landing Miniature Aerial Vehicle, US Patent 6,502,787,22. Barrett, R.M. 2004. Adaptive Aerostructures: The First Decade of Flight on Uninhabited Aerial Vehicles, In: Proceedings of SPIE, Smart Structures and Materials 2004: Industrial and Commercial Applications of Smart Structures Technologies, 36 March, San Diego, CA, Vol. 5388, pp. 190201. Barrett, R.M., Brozoski, F. and Gross, R.S. 1997. Design and Testing of a Subsonic All-Moving Adaptive Flight Control Surface, AIAA Journal, 35:12171219. Barrett, R.M., Burger, C. and Melian, J.P. 2001. Recent Advances in Uninhabited Aerial Vehicle (UAV) Flight Control with Adaptive Aerostructures, In: Proceedings of 4th European Demonstrators Conference, Edinburgh, Scotland. Barrett, R.M., Frye, P. and Schliesman, M. 1998. Design, Construction and Characterization of a Flightworthy Piezoelectric Solid State Adaptive Rotor, Journal of Smart Materials and Structures, 7:422431. Barrett, R.M. and Howard, N. 2000. Adaptive Aerostructures for Subscale Aircraft, In: Proceedings of 20th Southeastern Conference on Theoretical and Applied Mechanics, 1618 April, Pine Mountain, GA. Barrett, R.M. and Law, D. 1998. Design, Fabrication and Testing of a New Twist-Active Wing Design, In: Regelbrugge, M.E. (ed.), Proceedings of SPIE Smart Structures and Integrated Systems, 27 July, San Diego, CA, Vol. 3329, pp. 3341. Barrett, R.M. and Lee, G. 2000. Design Criteria, Aircraft Design, Fabrication and Testing of Sub-Canopy and Urban Micro-Aerial Vehicles, In: Proceedings of AIAA/AHS International Powered Lift Conference, November, Alexandria, VA.

Morphing Aircraft
Using an Inflatable Telescopic Spar, In: Proceedings of 44th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 710 April, Norfolk, VA, AIAA 20031718. Boeing, Aerospace Company. 1973. Variable Camber Wing, Second Progress Report. Bolonkin, A. and Gilyard, G.B. 1999. Estimated Benefits of VariableGeometry Wing Camber Control for Transport Aircraft, Technical Memorandum, NASA Dryden Flight Research Center, Edwards, CA, NASA/TM-1999-206586. Bonnema, K.L. and Smith, S.B. 1988. AFTI/F-111 Mission Adaptive Wing Flight Research Program, In: Proceedings of 4th AIAA Flight Test Conference, 1820 May, San Diego, CA, pp. 155161. Boothe, K. 2004. Dynamic Modeling and Flight Control of Morphing Air Vehicles, Masters Thesis, University of Florida, FL. Boria, F., Stanford, B., Bowman, S. and Ifju, P. 2009. Evolutionary Optimization of a Morphing Wing with Wind-Tunnel Hardware in the Loop, AIAA Journal, 47:399409. Bornengo, D., Scarpa, F. and Remillat, C. 2005. Evaluation of Hexagonal Chiral Structure for Morphing Airfoil Concept, Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering 219:185192. Bourdin, P., Gatto, A. and Friswell, M.I. 2007. Potential of Articulated Split Wingtips for Morphing-Based Control of Flying Wing, In: Proceedings of 25th AIAA Applied Aerodynamics Conference, 2528 June, Miami, Florida, AIAA 2007-4443 (16 pp). Bourdin, P., Gatto, A. and Friswell, M.I. 2008. Aircraft Control via Variable Cant-Angle Winglets, Journal of Aircraft, 45:414423. Bourdin, P., Gatto, A. and Friswell, M.I. 2010. Performing Co-ordinated Turns with Articulated Wing Tips as a Multi-axis Control Effectors, Aeronautical Journal, 114:3547. Bovais, C.S. and Davidson, P.T. 1994. Flight Testing the Flying Radar Target (FLYRT), In: Proceedings of 7th Biennial Flight Test Conference, 2023 June, Colorado Springs, CO, AIAA 942144, pp. 279286. Bowman, J., Sanders, B., Cannon, B., Kudva, J., Joshi, S. and Weisshaar, T. 2007. Development of Next Generation Morphing Aircraft Structures, In: Proceedings of 48th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 20071730 (10 pp). Bowman, J., Sanders, B. and Weisshaar, T. 2002. Evaluating the Impact of Morphing Technologies on Aircraft Performance, In: Proceedings of 43rd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2225 April, Denver, CO, AIAA 2002-1631 (11 pp). Boyd Jr. and Douglas, D. 2009. Aerodynamic and Acoustic Study of an Active Twist Rotor using a Loosely Coupled CFD/CSD Method, In: Proceedings of 35th European Rotorcraft Forum, 2225 September, Hamburg, GermanyArticle 1160. Brailovski, V., Terriault, P., Coutu, D., Georges, T., Morellon, E., rube , S. 2008. Morphing Laminar Wing Fischer, C. and Be with Flexible Extrados Powered by Shape Memory Alloy Actuators, In: Proceedings of International Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS 2008), 2830 October, Ellicott City, MDPaper SMASIS2008-377 (9 pp). Brender, S., Mark, H. and Aguilera, F. 1997. The Attributes of a Variable-Diameter Rotor System Applied to Civil Tiltrotor Aircraft, NASA Contractor Report, Ames Research Center, NASA-CR-203092. Bubert, E.A., Woods, B.K.S., Lee, K., Kothera, C.S. and Wereley, N.M. 2010. Design and Fabrication of a Passive 1D Morphing Aircraft Skin, Journal of Intelligent Material Systems and Structures, 21:16991717. Bushnell, G.S., Arbogast, D.J. and Ruggeri, R.T. 2008. Shape Control of a Morphing Structure (Rotor Blade) Using a Shape

869

Memory Alloy Actuator System, Future of SMA, In: Proceedings of SPIE Active and Passive Smart Structures and Integrated Systems 2008, 913 March, San Diego, CA, Vol. 6928, pp. 69282A69282A-11. Butt, L., Bilgen, O., Day, S., Sossi, C., Weaver, J., Wolek, A., Inman, D.J. and Mason, W.H. 2010b. Wing Morphing Design Utilizing Macro Fiber Composite, In: Proceedings of Society of Allied Weight Engineers (SAWE) 69th Annual Conference, May, Virginia Beach, VAPaper Number 3515-S. Butt, L., Day, S., Sossi, C., Weaver, J., Wolek, A., Bilgen, O., Inman, D.J. and Mason, W.H. 2010a. Wing Morphing Design Team Final Report 2010, Virginia Tech Departments of Mechanical Engineering and Aerospace and Ocean Engineering Senior Design Project, Blacksburg, VA. Bye, D.R. and McClure, P.D. 2007. Design of a Morphing Vehicle, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-1728 (16 pp). Caldwell, N., Gutmark, E. and Ruggeri, R. 2007. Performance Predictions of a Blade Twist Actuator System, In: Proceedings of 45th AIAA, Aerospace Sciences Meeting and Exhibit (2007), 811 January, Reno, NV, AIAA 2007-1216 (10 pp). Calkins, F.T., Mabe, J.H. and Ruggeri, R.T. 2008. Overview of Boeings Shape Memory Alloy Based Morphing Aerostructures, In: Proceedings of ASME Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS 2008), 2830 October, Ellicott City, MD, Paper SMASIS2008-648. Campanile, L.F. 2008. Modal Synthesis of Flexible Mechanisms for Airfoil Shape Control, Journal of Intelligent Material Systems and Structures, 19:779789. Campanile, L.F. and Sachau, D. 2000. The Belt-Rib Concept: A Structronic Approach to Variable Camber, Journal of Intelligent Material Systems and Structures, 11:215224. Cesnik, C.E.S. and Brown, E.L. 2003. Active Warping Control of a Joined-Wing Airplane Configuration, In: Proceedings of 44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, 710 April, Norfolk, VA, AIAA 2003-1715 (15 pp). Cesnik, C.E.S. and Shin, S.J. 2001. On the Twist Performance of a Multiple-Cell Active Helicopter Blade, Smart Materials and Structures, 10:5361. Cesnik, C.E.S., Shin, S.J. and Wilbur, M.L. 2001. Dynamic Response of Active Twist Rotor Blades, Smart Materials and Structures, 10:6276. Chandra, R. 2001. Active Shape Control of Composite Blades using Shape Memory Actuation, Smart Materials and Structures, 10:10181024. Chen, P.C., Sarhaddi, D., Jha, R., Liu, D.D., Griffin, K. and Yurkovich, R. 2000. Variable Stiffness Spar Approach for Aircraft Maneuver Enhancement Using Astros, Journal of Aircraft, 37:865871. Chopra, I. 2001. Recent Progress on Development of a Smart Rotor System, Presentation, NASA Langley Research Center, Langley, VA. Chopra, I. 2002. Review of State of Art of Smart Structures and Integrated Systems, AIAA Journal, 40:21452187. Chou, A. and Philen, M. 2008. High-Performance Flexible Matrix Composite Actuators for Trailing Edge Flap Control, In: Proceedings of Virginia Space Grant Consortium 2008 Student Research Conference, 21 April, Norfolk, VA (8 pp). Clarke, R., Allen, M.J., Dibley, R.P., Gera, J. and Hodgkinson, J. 2005. Flight Test of the F/A-18 Active Aeroelastic Wing Airplane, In: Proceedings of AIAA Atmospheric Flight Mechanics Conference and Exhibit, 1518 August, San Francisco, CA, AIAA 2005-6316 (31 pp). Cooper, J.E. 2006. Adaptive Stiffness Structures for Air Vehicle Drag Reduction, In: Multifunctional Structures/Integration of Sensors and Antennas, Meeting Proceedings RTO-MP-AVT-141, Neuillysur-Seine, France, pp. 15.115.12.

870

S. BARBARINO ET AL.
Florance, J.P., Burner, A.W., Fleming, G.A., Hunter, C.A. and Graves, S.S. 2003. Contributions of the NASA Langley Research Center to the DARPA/AFRL/NASA/Northrop Grumman Smart Wing Program, In: Proceedings of 44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics and Materials Conference, 710 April, Norfolk, VA, AIAA 20031961. Florance, J.R., Heeg, J., Spain, C.V. and Lively, P.S. 2004. Variable Stiffness Spar Wind-Tunnel Modal Development and Testing, In: Proceedings of 45th AIAA/ASME/ASCE/ AHS/ASC/ Structures, Structural Dynamics and Materials Conference, 1922 April, Palm Springs, CA, AIAA 2004-1588 (11 pp). Fradenburgh, E. 1973. Dynamic Model Wind Tunnel Tests of a Variable-Diameter, Telescoping-Blade Rotor System (TRAC Rotor), USAAMRDL Technical Report 73-32, Eustis Directorate, US Army Air Mobility R&D Laboratory, Fort Eustis, VA. Fradenburgh, E. and Matuska, D. 1992. Advancing Tiltrotor Stateof-the-Art with Variable Diameter Rotors, In: Proceedings of American Helicopter Society 48th Annual Forum, 35 June, Washington, DC. Frank, G.J., Joo, J.J., Sanders, B., Garner, D.M. and Murray, A.P. 2008. Mechanization of a High Aspect Ratio Wing for Aerodynamic Control, Journal of Intelligent Material Systems and Structures, 19:11021112. Frint, H., 1977. Design Selection Tests for TRAC Retraction Mechanism, USAAMRDL Technical Report 76-43, Eustis Directorate, US Army Air Mobility R&D Laboratory, Fort Eustis, VA. Gamboa, P., Aleixo, P., Vale, J., Lau, F. and Suleman, A. 2007. Design and Testing of a Morphing Wing for an Experimental UAV, RTO-MP-AVT-146, Neuilly-sur-Seine, France. Gamboa, P., Vale, J., Lau, F.J.P. and Suleman, A. 2009. Optimization of a Morphing Wing Based on Coupled Aerodynamic and Structural Constraints, AIAA Journal, 47:20872104. Gandhi, F. and Anusonti-Inthra, P. 2008. Skin Design Studies for Variable Camber Morphing Airfoils, Smart Materials and Structures, 17:015025. Gandhi, F., Frecker, M. and Nissly, A. 2008. Design Optimization of a Controllable Camber Rotor Airfoil, AIAA Journal, 46:142153. Gano, S.E., Perez, V.M., Renaud, J.E., Batill, S.M. and Sanders, B. 2004. Multilevel Variable Fidelity Optimization of a Morphing Unmanned Aerial Vehicle, In: Proceedings of 45th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 1922 April, Palm Springs, CA, AIAA 2004-1763 (16 pp). Gano, S.E. and Renaud, J.E. 2002. Optimized Unmanned Aerial Vehicle with Wing Morphing for Extended Range and Endurance, In: Proceedings of 9th AIAA/ISSMO Symposium and Exhibit on Multidisciplinary Analysis and Optimization, 46 September, Atlanta, GA, AIAA 2002-5668 (9 pp). Garcia, H.M., Abdulrahim, M. and Lind, R. 2003. Roll Control for a Micro Air Vehicle Using Active Wing Morphing, In: Proceedings of AIAA Guidance, Navigation, and Control Conference Exhibit, 1114 August, Austin, TX, AIAA J. 20035347 (10 pp). Garcia, H.M., Abdulrahim, M. and Lind, R. 2005. Flight Characteristics of Shaping the Membrane Wing of a Micro Air Vehicle, Journal of Aircraft, 42:131137. Gatto, A., Bourdin, P. and Friswell, M.I. 2010. Experimental Investigation into Articulated Winglet Effects on Flying Wing Surface Pressure Aerodynamics, Journal of Aircraft, 47:18111815. Geissler, W., Sobieczky, H. and Trenker, M. 2000. New Rotor Airfoil Design Procedure for Unsteady Flow Control, In: Proceedings of 26th European Rotorcraft Forum, 24 May, The Hague, HollandPaper No. 31.

Cooper, J.E., Amprikidis, M., Ameduri, S., Concilio, A., San Milan, J. and Castanon, M. 2005. Adaptive Stiffness Systems for an Active All-Moving Vertical Tail, In: Proceedings of European Conference for Aerospace Sciences (EUCASS), 47 July, Moscow, Russia (7 pp). Crawley, E.F., Lazarus, K.B. and Warkentin, D.J. 1989. Embedded Actuation and Processing in Intelligent Materials, In: Proceedings of 2nd International Workshop on Composite Materials and Structures, 1415 September, Troy, NY. Cuji, E. and Garcia, E. 2008. Aircraft Dynamics for Symmetric and Asymmetric V-Shape Morphing Wings, In: ASME Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS08), 2830 October, Ellicott City, MDSMASIS 2008-424. DARPA Tactical Technology Office (TTO). 2011. Available at: http:// www.darpa.mil/Our_Work/TTO/Programs/DiscRotor/DiscRotor_Compound_Helicopter.aspx (accessed date May 22, 2011). Davis, S.J., Moffitt, R., Quackenbush, T.R. and Wachspress, D.A. 1995. Aerodynamic Design Optimization of a Variable Diameter Tilt Rotor, In: Proceedings of American Helicopter Society 51st Annual Forum, 911 May 1995, Fort Worth, TX, pp. 101111. Daynes, S., Nall, S.J., Weaver, P.M., Potter, K.D., Margaris, P. and Mellor, P.H. 2009. On a Bistable Flap for an Airfoil, In: Proceedings of 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2103. de Camp, R.W. and Hardy, R. 1984. Mission Adaptive Wing Advanced Research Concepts, AIAA 1984-2088. Derham, R.C. and Hagood, N.W. 1996. Rotor Design Using Smart Materials to Actively Twist Blades, In: Proceedings of American Helicopter Society 52nd Annual Forum, June, Washington, DC, Vol. 2, pp. 12421252. Detrick, M. and Washington, G. 2007. Modeling and Design of a Morphing Wing for Micro Unmanned Aerial Vehicles via Active Twist, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Material Conference, 2326 April, Honolulu, HI, AIAA 2007-1788 (7 pp). Diaconu, C.G., Weaver, P.M. and Mattioni, F. 2008. Concepts for Morphing Airfoil Sections Using Bi-Stable Laminated Composite Structures, Thin-Walled Structures, 46:689701. Edi, P. and Fielding, J.P. 2006. Civil-Transport Wing Design Concept Exploiting New Technologies, Journal of Aircraft, 43:932940. Eggleston, G., Hutchison, C., Johnston, C., Koch, B., Wargo, G. and Williams, K. 2002. Morphing Aircraft Design Team, Virginia Tech Aerospace Engineering Senior Design Project. Ehlers, S.M. and Weisshaar, T.A. 1990. Static Aeroelastic Behavior of an Adaptive Laminated Piezoelectric Composite Wing, AIAA Journal, 28:16111623. Ehlers, S.M. and Weisshaar, T.A. 1992. Adaptive Wing Flexural Axis Control, In: Proceedings of Third International Conference on Adaptive Structures, 911 November, San Diego, CA. Elzey, D.M., Sofla, A.Y.N. and Wadley, H.N.G. 2003. A BioInspired, High-Authority Actuator for Shape Morphing Structures, In: Logoudas, D.C. (ed.), Proceedings of SPIE, Smart Structures and Materials 2003: Active Materials: Behavior and Mechanics, 13 August, San Diego, CA, Vol. 5053, pp. 92100. Epps, J.J. and Chopra, I. 2001. In-Flight Tracking of Helicopter Rotor Blades Using Shape Memory Alloy Actuators, Smart Materials and Structures, 10:104111. Flanagan, J.S., Strutzenberg, R.C., Myers, R.B. and Rodrian, J.E. 2007. Development and Flight Testing of a Morphing Aircraft, the NextGen MFX-1, In: Proceedings of 48th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics, and Materials, 2326 April, Honolulu, HI, AIAA 2007-1707. FlexSys Inc. 2011. Available at: http://www.flxsys.com/index.shtml (accessed date May 22, 2011).

Morphing Aircraft
Gern, F.H., Inman, D.J. and Kapania, R.K. 2005. Computation of Actuation Power Requirements for Smart Wings with Morphing Airfoils, AIAA Journal, 43:24812486. Gevers Aircraft Inc. 1997. Multi-Purpose Aircraft, United States Patents Office, Patent Number 5,645,250. Giurgiutiu, V. 2000. Review of Smart-Materials Actuation Solutions for Aeroelastic and Vibration Control, Journal of Intelligent Material Systems and Structures, 11:525544. Giurgiutiu, V., Chaudhry, Z. and Rogers, C.A. 1994a. Active Control of Helicopter Rotor Blades with Induced Strain Actuators, In: Proceedings of 35th AIAA/ASME/ASCE/AHS/ ASC Structures, Structural Dynamics, and Materials Conference, 1820 April, Hilton Head, SC (10 pp). Giurgiutiu, V., Chaudhry, Z. and Rogers, C.A. 1994b. Engineering Feasibility of Induced-Strain Actuators for Rotor Blade Active Vibration Control, In: Proceedings of SPIE Smart Structures and Materials 94 Conference, 1318 February, Orlando, FL, Vol. 2190, Paper # 2190-11, pp. 107122. Grady, B. 2010. Multi-Objective Optimization of a Three Cell Morphing Wing Substructure, Masters Thesis, University of Dayton, OH. Grant, D., Abdulrahim, M. and Lind, R. 2006. Flight Dynamics of a Morphing Aircraft Utilizing Independent Multiple-Joint Wing Sweep, In: Proceedings of AIAA Atmospheric Flight Mechanics Conference and Exhibit, 2124 August, Keystone, CO, AIAA 2006-6505 (15 pp). Griffin, K.E. and Hopkins, M.A. 1997. Smart Stiffness for Improved Roll Control, Journal of Aircraft, Engineering Notes, 34:445447. Grohmann, B.A., Maucher, C.K. and Ja nker, P. 2006. Actuation Concepts for Morphing Helicopter Rotor Blades, In: Proceedings of 25th International Congress of the Aeronautical Sciences, 38 September, Hamburg, Germany (10 pp). Grohmann, B.A., Maucher, C.K., Janker, P. and Wierach, P. 2008. Embedded Piezoceramic Actuators for Smart Helicopter Rotor Blades, In: 49th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials, 710 April, Schaumburg, IL, AIAA 2008-1702 (10 pp). Guiler, R. and Huebsch, W. 2005. Wind Tunnel Analysis of a Morphing Swept Tailless Aircraft, In: Proceedings of 23rd AIAA Applied Aerodynamics Conference, 69 June, Toronto, Ontario, Canada, AIAA 2005-4981 (14 pp). Hall, J., Mohseni, K. and Lawrence, D. 2005. Investigation of Variable Wing-Sweep for Applications in Micro Air Vehicles, In: Proceedings of Infotech@Aerospace, 2629 September, Arlington, VA, AIAA 2005-7171 (11 pp). Han, C., Ruy, K. and Lee, S. 2007. Experimental Study of a Telescopic Wing Inside Channel, Engineering Notes, Journal of Aircraft, 44:10291030. Henry, J.J. and Pines, D.J. 2007. A Mathematical Model for Roll Dynamics by Use of a Morphing-Span Wing, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-1708 (16 pp). Heryawan, Y., Park, H.C., Goo, N.S., Yoon, K.J. and Byun, Y.H. 2005. Design and Demonstration of a Small Expandable Morphing Wing, In: Proceedings of SPIE 12th Annual International Symposium on Smart Structures and Materials, 610 March, San Diego, CA, 5764: 224-231. Hu, H., Tamai, M. and Murphy, J. 2008. Flexible Membrane Airfoils at Low Reynolds Numbers, Journal of Aircraft, 45:17671778. Hunton, L.W. and Dew, J.K. 1947. Measurements of the Damping in Roll of Large-Scale swept-Forward and SweptBack Wings, Ames Aeronautical Laboratory, NACA RMA7D11. Icardi, U. and Ferrero, L. 2009. Preliminary Study of an Adaptive Wing with Shape Memory Alloy Torsion Actuators, Materials and Design, 30:42004210.

871

Ifju, P.G., Ettinger, S., Jenkins, D.A. and Martinez, L. 2001. Composite Materials for Micro Air Vehicles, In: Proceedings of SAMPE Annual Conference, 610 May, Long Beach, CA (12 pp). Ifju, P.G., Jenkins, D.A., Ettinger, S., Lian, Y., Shyy, W. and Waszak, M.R. 2002. Flexible-Wing-Based Micro Air Vehicles, In: Proceedings of AIAA Annual Conference, January, Reno, NV, AIAA 2002-0705 (13 pp). Inman, D.J. 2001. Wings: Out of the Box. Determining Actuator Requirements for Controlled Morphing Air Vehicles Aerodinamic Loads, DARPA Technology Interchange Meeting, Wright Patterson Air Force Base, Dayton, OH. Ivanco, T.G., Scott, R.C., Love, M.H., Zink, S. and Weisshaar, T.A. 2007. Validation of the Lockheed Martin Morphing Concept with Wind Tunnel Testing, In: Proceedings of 48th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 20072235 (17 pp). Janker, P., Kloppel, V., Hermle, F., Lorkowski, T., Storm, S., Christmann, M. and Wettemann, M. 1999. Development and Evaluation of a Hybrid Piezoelectric Actuator for Advanced Flap Control Technology, In: Proceedings of 25th European Rotorcraft Forum (ERF 1999), 1416 September, Rome, Italy. Jha, A.K. and Kudva, J.N. 2004. Morphing Aircraft Concepts, Classifications, and Challenges, In: Anderson, E.H. (ed.), Proceedings of SPIE Smart Structures and Materials 2004: Industrial and Commercial Applications, 29 July, San Diego, CA, Vol. 5388, pp. 213224. Johnson, T., Frecker, M., Abdalla, M., Gurdal, Z. and Lindner, D. 2009. Nonlinear Analysis and Optimization of Diamond Cell Morphing Wings, Journal of Intelligent Material Systems and Structures, 20:815824. Johnston, C.O., Mason, W.H., Han, C. and Inman, D.J. 2007. Actuator-Work Concepts Applied to Unconventional Aerodynamic Control Devices, Journal of Aircraft, 44:14591468. Johnston, C.O., Neal, D.A., Wiggins, L.D., Robertshaw, H.H., Mason, W.H. and Inman, D.J. 2003. A Model to Compare the Flight Control Energy Requirements of Morphing and Conventionally Actuated Wings, In: Proceedings of 44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, 710 April, Norfolk, VA, AIAA 20031716 (9 pp). Jones, R.T. 1945. Wing Plan Forms for High-Speed Flight, NACA Report 863. Joo, J.J., Reich, G.W. and Westfall, J.T. 2009. Flexible Skin Development for Morphing Aircraft Applications Via Topology Optimization, Journal of Intelligent Material Systems and Structures, 20:19691985. Joo, J.J., Sanders, B., Johnson, T. and Frecker, M.I. 2006. Optimal Actuator Location within a Morphing Wing Scissor Mechanism Configuration, In: Proceedings of SPIE Smart Structures and Materials 2006: Modeling, Signal Processing, and Control, 27 February2 March, San Diego, CA, Vol. 6166 (12 pp). Joshi, S.P., Tidwell, Z., Crossley, W.A. and Ramakrishnan, S. 2004. Comparison of Morphing Wing Strategies Based Upon Aircraft Performance Impacts, In: Proceedings of 45th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, Palm Springs, CA, AIAA 2004-1722 (7 pp). Kerho, M. 2007. Adaptive Airfoil Dynamic Stall Control, Journal of Aircraft, 44:13501360. Khoshlahjeh, M., Bae, E.S. and Gandhi, F. 2010. Helicopter Performance Improvement with Variable Chord Morphing Rotors, In: Proceedings of 36th European Rotorcraft Forum, 79 September, Paris, France (12 pp). Khoshlahjeh, M., Gandhi, F. and Webster, S. 2011. Extendable Chord Rotors for Helicopter Envelope Expansion and Performance Improvement, In: Proceedings of 67th American Helicopter Society (AHS) Annual Forum, 35 May, Virginia Beach, VA.

872

S. BARBARINO ET AL.
Lee, T. 1999. Design of High Displacement Smart Trailing Edge Flap Actuator Incorporating Dual-Stage Mechanical Stroke Amplifiers for Rotors, Ph.D. Dissertation, Dept. of Aerospace Engineering, University of Maryland, College Park, December 1999. Lee, T. and Chopra, I. 2001a. Design of Piezostack-Driven TrailingEdge Flap Actuator for Helicopter Rotors, Smart Materials and Structures, 10:1524. Lee, T. and Chopra, I. 2001b. Wind Tunnel Test of Blade Sections with Piezoelectric Trailing-Edge Flap Mechanism, In: Proceedings of 57th Annual Forum of the American Helicopter Society, 911 May, Washington, DC, Vol. 2, pp. 19121923. Leishman, G.J. 2006. Principles of Helicopter Aerodynamics, Cambridge Aerospace Series No. 12, Cambridge University Press, ISBN: 0521858607. Leite, A., Vale, J., Lau, F. and Suleman, A. 2009. Development of Morphing Strategies for Flight Demonstrator RPV, In: Proceedings of 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2134 (15 pp). Leon, O. and Gandhi, F. 2009. Rotor Power Reduction using Multiple Spanwise-Segmented Optimally-Actuated TrailingEdge Flaps, In: Proceedings of 35th European Rotorcraft Forum, 2225 September, Hamburg, Germany, Article 1309. Leon, O., Hayden, E. and Gandhi, F. 2009. Rotorcraft Operating Envelope Expansion Using Extendable Chord Sections, In: Proceedings of American Helicopter Society 65th Annual Forum, 2729 May, Grapevine, TX (14 pp). Lesieutre, G. and Davis, C. 1997. Can a Coupling Coefficient of a Piezoelectric Device be Higher Than Those of Its Active Material? Journal of Intelligent Materials Systems and Structures, 8:859867. Loftin, L. 1985. Quest for Performance, NASA SP 468. Love, M.H., Zink, P.S., Stroud, R.L., Bye, D.R., Rizk, S. and White, D. 2007. Demonstration of Morphing Technology through Ground and Wind Tunnel Tests, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-1729 (12 pp). Lv, H., Leng, J. and Du, S. 2009. A Survey of Adaptive Materials and Structures Research in China, In: Proceedings of 50th AIAA/ASME/ASCE/ASH/ACS Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2389 (8 pp). Madden, J.D.W., Schmid, B., Hechinger, M., Lafontaine, S.R., Madden, P.G.A. and Hover, F.S. 2004. Application of Polypyrrole Actuators: Feasibility of variable camber Foils, IEEE Journal of Oceanic Engineering, 29:738749. Majji, M., Rediniotis, O. and Junkins, J.L. 2007. Design of a Morphing Wing: Modeling and Experiments, In: Proceedings of AIAA Atmospheric Flight Mechanics Conference and Exhibit, 2023 August, Hilton Head, SC, AIAA 2007-6310 (9 pp). Manzo, J. and Garcia, E. 2010. Demonstration of an in situ Morphing Hyperelliptical Cambered Span Wing Mechanism, Smart Materials and Structures, 19:025012. Manzo, J., Garcia, E., Wickenheiser, A. and Horner, G.C. 2005. Design of a Shape-Memory Alloy Actuated Macro-scale Morphing Aircraft Mechanism, In: Flatau, A.B. (ed.), Proceedings of SPIE Smart Structures and Materials 2005: Smart Structures and Integrated Systems, 17 May, San Diego, CA, Vol. 5764, p. 744. Marmier, P. and Wereley, N.M. 2003. Control of Sweep Using Pneumatic Actuators to Morph Wings of Small Scale UAVs, In: Proceedings of 44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, 710 April, Norfolk, VA, AIAA 2003-1802 (10 pp). Marques, M., Gamboa, P. and Andrade, E. 2009. Design and Testing of a Variable Camber Flap for Improved Efficiency, In: Proceedings of RTO Applied Vehicle Technology Panel (AVT) Symposium, 2024 April, Evora, Portugal (22 pp).

Kim, D.K. and Han, J.H. 2006. Smart Flapping Wing using MacroFiber Composite Actuators, In: Matsuzaki, Y. (ed.), Proceedings of SPIE Smart Structures and Materials 2006: Smart Structures and Integrated Systems, San Diego, CA, Vol. 6173, pp. 133141. Kim, D.K., Han, J.H. and Kwon, K.J. 2009. Wind Tunnel Tests for a Flapping Wing Model with a Changeable Camber using MacroFiber Composite Actuators, Smart Materials and Structures, 18:024008. Koratkar, N.A. and Chopra, I. 2000. Development of a Mach-Scaled Model with Piezoelectric Bender Actuated Trailing-Edge Flaps for Helicopter Individual Blade Control (IBC), AIAA Journal, 38:11131124. Koratkar, N.A. and Chopra, I. 2001. Wind Tunnel Testing of a Mach-Scaled Rotor Model with Trailing-Edge Flaps, Smart Materials and Structures, 10:114. Kota, S., Ervin, G., Osborn, R. and Ormiston, R.A. 2008. Design and Fabrication of an Adaptive Leading Edge Rotor Blade, In: Proceedings of American Helicopter Society 64th Annual Forum, 29 April1 May, Montreal, Canada (9 pp). Kota, S., Osborn, R., Ervin, G., Maric, D., Flick, P. and Paul, D. 2009. Mission Adaptive Compliant Wing - Design, Fabrication and Flight Test, In: RTO Applied Vehicle Technology Panel (AVT) Symposium, Evora, Portugal, RTO-MP-AVT-168. Kovalovs, A., Barkanov, E. and Gluhihs, S. 2007. Active Twist of Model Rotor Blades with D-Spar Design, Transport, 22:3844. Kress, R.W. 1983. Variable Sweep Wing Design, In: Proceedings of Aircraft Prototype and Technology Demonstrator Symposium, 2324 March, Dayton, OH, AIAA 1983-1051. Kudva, J.N. 2001. Overview of the DARPA/AFRL/NASA Smart Wing Phase II Program, In: Proceedings of SPIE Smart Structures and Materials Conference, Vol. 4332, pp. 383389. Kudva, J.N. 2004. Overview of the DARPA Smart Wing Project, Journal of Intelligent Material Systems and Structures, 15:261267. Kudva, J.N. and Carpenter, B. 2000. Smart Wing Program, In: Proceedings of DARPA Technology Interchange Meeting, June. Kudva, J.N., Jardine, P., Martin, C. and Appa, K. 1996a. Overview of the ARPA/WL Smart Structures and Material Development Smart Wing Contract, In: Proceedings of SPIE North American Conference on Smart Structures and Materials, San Diego, CASPIE, 2721: 1016. Kudva, J.N., Lockyer, A.J. and Appa, K. 1996b. Adaptive Aircraft Wing, In: Proceedings of AGARD SMP Lecture series 205, Smart Structures and Materials: Implications for Military Aircraft of New Generation, 3031 October, Philadelphia, PAAGARD-LS-205. Kuzmina, S., Gennadi, A., Schweiger, J., Cooper, J., Amprikidis, M. and Sensberg, O. 2002. Review and Outlook on Active and Passive Aeroelastic Design Concepts for Future Aircraft, In: Proceedings of International Congress of the Aeronautical Sciences (ICAS 2002), pp. 432.1432.10. Lajux, V. and Fielding, J. 2005. Development of a Variable Camber Leading Edge Device Design Methodology, In: Proceedings of AIAA 5th ATIO and 16th Lighter-Than-Air Sys Tech. and Balloon Systems Conferences, 28 September, Arlington, VA, AIAA 20057356. Lazarus, K.B., Crawley, E.F. and Bohlmann, J.D. 1991. Static Aeroelastic Control Using Strain Actuated Adaptive Structures, Journal of Intelligent Material Systems and Structures, 2:386410. Lazos, B.S. and Visser, K.D. 2006. Aerodynamic Comparison of Hyper-Elliptic Cambered Span (HECS) Wings with Conventional Configurations, In: Proceedings of 24th AIAA Applied Aerodynamic Conference, 58 June, San Francisco, CA, AIAA 2006-3469 (18 pp). Lee, G. 1997. Design and Testing of the Kolibri Vertical Take-Off and Landing Micro Aerial Vehicle, Final Report for the Department of Defense CounterDrug Technology Office.

Morphing Aircraft
Martin, C.A., Bartley-Cho, J., Flanagan, J.S. and Carpenter, B.F. 1999. Design and Fabrication of Smart Wing Wind Tunnel Model and SMA Control Surfaces, In: Jacobs, J.H. (ed.), Proceedings of SPIE Conference on Industrial and Commercial Applications of Smart Structures Technologies, Newport Beach, CA, Vol. 3674, pp. 237248. Martin, S. and Hall, G. 1969. Rotor Will Retract in Flight to Improve V/STOL Cruise Performance, Space Aeronautics. Martin, C.A., Hallam, B.J., Flanagan, J.S. and Bartley-Cho, J. 2002. Design, Fabrication and Testing of Scaled Wind Tunnel Model for the Smart Wing Phase 2 Program, In: McGowan, A.R. (ed.), Proceedings of SPIE Smart Structures and Materials 2002: Industrial and Commercial Applications of Smart Structures Technologies, San Diego, CA, Vol. 4698, pp. 4452. Martin, C.A., Kudva, J., Austin, F., Jardine, A.P., Scherer, L.B., Lockyer, A.J. and Carpenter, B.F. 1998. Smart Materials and Structures-Smart Wing Volumes I, II, III, and IV, Report: AFRL-ML-WP-TR-1999-4162, Northrop Grumman Corporation, Hawthorne, CA. Martins, A.L. and Catalano, F.M. 1998. Viscous Drag Optimization for a Transport Aircraft Mission Adaptive Wing, In: Proceedings of 21st ICAS Congress, 1318 September, Melbourne, AustraliaPaper A98-31499. Mattioni, F., Gatto, A., Weaver, P.M., Friswell, M.I. and Potter, K.D. 2006. The Application of Residual Stress Tailoring of Snapthrough Composites for Variable Sweep Wings, In: Proceedings of 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 14 May, Newport, RI, AIAA 2006-1972. McCormik, B.W. 1995. Aerodynamics, Aeronautics and Flight Mechanics, 2 edn, Wiley, New York. McGowan, A.R. 2008. Overview: Morphing Activities in the USA, In: Proceedings of Advanced Course on Morphing Aircraft Materials, Mechanisms and Systems, 1720 November, Lisbon, Portugal. McGowan, A.R., Horta, L.G., Harrison, J.S. and Raney, D.L. 1999. Research Activities within NASAs Morphing Program, In: RTO Meeting Proceedings 36 (AC/323(AVT)TP/17) from the RTO AVT Specialists Meeting on Structural Aspects of Flexible Aircraft Control, October, Ottawa, AL, Canada. McKnight, G., Doty, R., Keefe, A., Herrera, G. and Henry, C. 2010. Segmented Reinforcement Variable Stiffness Materials for Reconfigurable Surfaces, Journal of Intelligent Material Systems and Structures, 21:17831793. Mesaric, M. and Kosel, F. 2004. Unsteady Airload of an Airfoil with Variable Camber, Aerospace Science and Technology, 8:167174. Miller, G.D. 1988. Active Flexible Wing (AFW) Technology, Report: AFWAL-TR-87-3096, Rockwell International North American Aircraft Operations, Los Angeles, CA. Miller, S., Vio, G.A., Cooper, J.E., Vale, J., da Luz, L., Gomes, A., Lau, F., Suleman, A., Cavagna, L., De Gaspari, A., Ricci, S., Riccobene, L., Scotti, A. and Terraneo, M. 2010. SMorph Smart Aircraft Morphing Technologies Project, In: Proceedings of 51st AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 1215 April, Orlando, FL, AIAA 2010-2742. Mirone, G. 2007. Design and Demonstrators Testing of Adaptive Airfoils and Hingeless Wings Actuated by Shape Memory Alloy Wires, Smart Structures and Systems, 3:89114. Mirone, G. and Pellegrino, A. 2009. Prototipo di Sezione Alare Adattiva per UAV: Centina Deformabile con Attuatori in SMA, In: Proceeding of the XXXVIII Convegno Nazionale, Associazione Italiana per lAnalisi delle Sollecitazioni (AIAS), 911 September, Turin, Italy, in Italian. Mistry, M., Gandhi, F. and Chandra, R. 2008. Twist Control of an IBeam Through Vlasov Bimoment Actuation, In: Proceedings of 49th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Material Conference, 711 April, Schaumburg, IL, AIAA 2008-2278 (12 pp).

873

Mistry, M., Nagelsmit, M., Ganhdi, F. and Gurdal, Z. 2010. Actuation Requirements of a Warp Induced Variable Twist Rotor Blade, In: Proceedings of AMSE 2010 Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS2010), 28 September1 October, Philadelphia, PASMASIS 2010-3615. Monner, H.P. 2001. Realization of an Optimized Wing Camber by Using Form Variable Flap Structures, Aerospace Science and Technology, 5:445455. Monner, H.P., Bein, T., Hanselka, H. and Breitbach, E. 1998. Design Aspects of the Adaptive Wing The Elastic Trailing Edge and the Local Spoiler Bump, Royal Aeronautical Society, Multidisciplinary Design and Optimization, London, UK. Monner, H.P., Kintscher, M., Lorkowski, T. and Storm, S. 2009. Design of a Smart Droop Nose as Leading Edge High Lift System for Transportation Aircrafts, In: Proceedings of 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 45 May, Palm Springs, CA, AIAA 2009-2128 (10 pp). Monner, H.P., Sachau, D. and Breitbach, E. 1999. Design Aspects of the Elastic Trailing Edge for an Adaptive Wing, In: Proceedings of RTO AVT Specialists Meeting on Structural Aspects of Flexible Aircraft Control, 1820 October, Ottawa, Canada, Vol. RTO MP-36 (8 pp). Moorhouse, D., Sanders, B., von Spakovsky, M. and Butt, J. 2006. Benefits and Design Challenges of Adaptive Structures for Morphing Aircraft, Aeronautical Journal, Royal Aeronautical Society, London, 110:157162. Munday, D. and Jacob, J. 2001. Active Control of Separation on a Wing with Conformal Camber, AIAA Journal, 20010293. Murray, G., Gandhi, F. and Bakis, C. 2010. Flexible Matrix Composite Skins for One-Dimensional Wing Morphing, Journal of Intelligent Material Systems and Structures, 21:17711781. Murua, J., Palacios, R. and Peiro, J. 2010. Camber Effects in the Dynamic Aeroelasticity of Compliant Airfoils, Journal of Fluids and Structures, 26:527543. Musgrove, R.G. 1976. Unites States Patent 3,944,170, Apparatus For Producing Pivotal Movement, 16 March 1976. Musgrove, R.G. 1981. Unites States Patent 4,286,761, Eccentric Actuator, 1 September 1981. Nam, C., Chattopadhyay, A. and Kim, Y. 2002. Application of Shape Memory Alloy (SMA) Spars for Aircraft Maneuver Enhancement, In: Davis, L.P. (ed.), Proceedings of SPIE Smart Structures and Materials 2002: Smart Structures and Integrated Systems, 15 July, San Diego, CA, Vol. 4701, pp. 226236. Nam, C., Chen, P.C., Sarhaddi, D., Liu, D., Griffin, K. and Yurkovich, R. 2000. Torsion-Free Wing Concept for Aircraft Maneuver Enhancement, In: Proceedings of 41st AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 36 April, Atlanta, GA, AIAA 2000-1620. Namgoong, H., Crossley, W.A. and Lyrintzis, A.S. 2006. Aerodynamic Optimization of a Morphing Airfoil Using Energy as an Objective, In: Proceedings of 44th AIAA Aerospace Sciences Meeting and Exhibit, 912 January, Reno, NV, AIAA 2006-1324. Neal, D.A., Farmer, J. and Inman, D.J. 2006. Development of a Morphing Aircraft Model for Wind Tunnel Experimentation, In: Proceedings of 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 14 May, Newport, RI, AIAA 2006-2141. Neal, D.A., Good, M.G., Johnston, C.O., Robertshaw, H.H., Mason, W.H. and Inman, D.J. 2004. Desing and Wind-Tunnel Analysis of a Fully Adaptive Aircraft Configuration, In: 45th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 1922 April, Palm Springs, CA, AIAA 2004-1727. Null, W. and Shkarayev, S. 2005. Effect of Camber on the Aerodynamics of Adaptive-Wing Micro Air Vehicles, Journal of Aircraft, 42:15371542.

874

S. BARBARINO ET AL.
Powers, S.G., Webb, L.D., Friend, E.L. and Lokos, W.A. 1992. Flight Test Results from a Supercritical Mission Adaptive Wing With Smooth Variable Camber, Technical Memorandum, NASA 4415. Prabhakar, T., Gandhi, F. and McLaughlin, D. 2007. A Centrifugal Force Actuated Variable Span Morphing Helicopter Rotor, In: Proceedings of 63rd Annual AHS International Forum and Technology Display, 13 May, Virginia Beach, VA. Prahlad, H. and Chopra, I. 2001. Design of a Variable Twist Tiltrotor Blade Using Shape Memory Alloy (SMA) Actuators, In: Prahlad, H. (ed.), Proceedings of SPIE Smart Structures and Materials 2001: Smart Structures and Integrated Systems, 16 August, Newport Beach, CA, Vol. 4327, pp. 4659. Prahlad, H. and Chopra, I. 2002. Characterization of SMA Torsional Actuators for Active Twist of Tilt Rotor Blades, In: Proceedings of 43rd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Material Conference, 2225 April, Denver, CO, AIAA 2002-1445. Prock, B.C., Weisshaar, T.A. and Crossley, W.A. 2002. Morphing Airfoil Shape Change Optimization With Minimum Actuator Energy As An Objective, In: Proceedings of 9th AIAA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, 46 September, Atlanta, GA, AIAA 2002-5401 (13 pp). Reed, J.L., Hemmelgarn, C.D., Pelley, B.M. and Havens, E. 2005. Adaptive Wing Structures, In: Proceedings of SPIE Smart Structures and Materials 2005: Industrial and Commercial Applications of Smart Structures Technologies, 9 March, San Diego, CA, Vol. 5762. Reich, G.W. and Sanders, B. 2007. Introduction to Morphing Aircraft Research, Journal of Aircraft, 44:1059. Renken, J.H. 1985. Mission Adaptive Wing Camber Control Systems for Transport Aircraft, In: Proceedings of 3rd Applied Aerodynamics Conference, 1416 October, Colorado Springs, CO , AIAA 85-5006. Ricci, S. 2008. Adaptive Camber Mechanism for Morphing Experiences at DIA-PoliMI, In: Proceedings of Advanced Course on Morphing Aircraft Materials, Mechanisms and Systems, 1720 November, Lisbon, Portugal. Rodriguez, A.R. 2007. Morphing Aircraft Technology Survey, In: Proceedings of 45th AIAA Aerospace Sciences Meeting and Exhibit, 811 January, Reno, NV, AIAA 2007-1258 (16 pp). Rogers, J.P. and Hagood, N.W. 1997. Design and Manufacture of an Integral Twist-Actuated Rotor Blade, In: Proceedings of 38th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Adaptive Structures Forum, Reston, VA. Roglin, R.L. and Hanagud, S.V. 1996. A Helicopter with Adaptive Rotor Blades for Collective Control, Smart Materials and Structures, 5:7688. Roglin, R.L., Hanagud, S.V. and Kondor, S. 1994. Adaptive Airfoils for Helicopters, In: Proceedings of 35th AIAA Structures, Structural Dynamics, and Materials Conference, 1820 April, Hilton Head, SC, pp. 279287. Roth, B., Peters, C. and Crossley, W. 2002. Aircraft Sizing with Morphing as an Independent Variable: Motivation, Strategies, and Investigations, In: Proceedings of AIAA Aircraft Technology, Integration, and Operations (ATIO) 2002 Conference, 13 October, Los Angeles, CA, AIAA 2002-5840. Ruggeri, R.T., Arbogast, D.J. and Bussom, R.C. 2008a. Wind Tunnel Testing of a Lightweight - Scale Actuator Utilizing Shape Memory Alloy, In: Proceedings of 49th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 710 April, Schaumburg, IL, AIAA 2008-2279. Ruggeri, R.T., Bussom, R.C. and Arbogast, D.J. 2008b. Development of a -scale NiTinol actuator for reconfigurable structures, In: Proceedings of SPIE Industrial and Commercial Applications of Smart Structures Technologies 2008, 10 March, San Diego, CA, Vol. 6930doi:10.1117/12.7759. Runge, J.B., Osmont, D. and Ohayon, R. 2010. Twist Control of Aerodynamic Profiles By a Reactive Method (Experimental

Olympio, K.R. and Gandhi, F. 2010a. Flexible Skins for Morphing Aircraft Using Cellular Honeycomb Cores, Journal of Intelligent Material Systems and Structures, 21:17191735. Olympio, K.R. and Gandhi, F. 2010b. Zero Poissons Ratio Cellular Honeycombs for Flex Skins Undergoing One-Dimensional Morphing, Journal of Intelligent Material Systems and Structures, 21:17371753. Olympio, K.R., Gandhi, F., Asheghian, L. and Kudva, J. 2010. Design of a Flexible Skin for a Shear Morphing Wing, Journal of Intelligent Material Systems and Structures, 21:17551770. Pagano, A., Ameduri, S., Cokonaj, V., Prachar, A., Zachariadis, Z. and Drikakis, D. 2009. Helicopter Blade Twist Control through SMA Technology: Optimal Shape Prediction, Twist Actuator Realisations and Main Rotor Enhanced Performance Computation, In: Proceedings of 35th European Rotorcraft Forum, 2225 September, Hamburg, Germany, Paper 1297 (11 pp). Paradies, R. and Ciresa, P. 2009. Active Wing Design with Integrated Flight Control Using Piezoelectric Macro Fiber Composites, Smart Material and Structure, 18:035010. Parker, H.F. 1920. The Parker Variable Camber Wing, Annual Report, National Advisory Committee for Aeronautics, Report No.77. Peel, L.D., Mejia, J., Narvaez, B., Thompson, K. and Lingala, M. 2008. Development of a Simple Morphing Wing Using Elastomeric Composites as Skins and Actuators, In: Proceedings of ASME Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS 2008), 2830 October, Ellicott City, MD, Paper SMASIS2008-544 (8 pp). Peel, L.D., Mejia, J., Narvaez, B., Thompson, K. and Lingala, M. 2009. Development of a Simple Morphing Wing Using Elastomeric Composites as Skins and Actuators, Journal of Mechanical Design, 131:091003 (8 pp). Pendleton, E. 2000. Back to the Future-How Active Aeroelastic Wings are a Return to Aviations Beginnings and a Small Step to Future Bird-Like Wings, In: Proceedings of RTO AVT Symposium on Active Control Technology for Enhanced Performance Operational Capabilities of Military Aircraft, Land Vehicles, and Sea Vehicles, 811 May, Braunschweig, Germany (8 pp). Pendleton, E.W., Bessette, D., Field, P.B., Miller, G.D. and Griffin, K.E. 2000. Active Aeroelastic Wing Flight Research Program: Technical Program and Model Analytical Development, Journal of Aircraft, 37:554561. Perera, M. and Guo, S. 2009. Optimal Design of a Seamless Aeroelastic Wing Structure, In: Proceedings of 50th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2195 (14 pp). Perkins, D.A., Reed, J.L. and Havens, E. 2004. Morphing Wing Structures for Loitering Air Vehicles, In: Proceedings of 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics & Materials Conference, 1922 April, Palm Springs, CA, AIAA 2004-1888 (10 pp). Perry, R.L. 1966. Variable Sweep: A Case History of Multiple Re-Innovation, p. 3459, The Rand Corporation, Santa Monica, CA. Pinkerton, J.L. and Moses, R.W. 1997. A Feasibility Study to Control Airfoil Shape Using THUNDER, Technical Memorandum, NASA Langley Research Center, Hampton, VA. Polhamus, E.C. and Hammond, A.D. 1967. Aerodynamic Research Relative to Variable-Sweep Multimission Aircraft, Langley Research Center, NASA N67-33072. Poonsong, P. 2004. Design and Analysis of a Multi-Section Variable Camber Wing, Masters Thesis, University of Maryland, MD. Popov, A.V., Labib, M., Fays, J. and Botez, R.M. 2008. ClosedLoop Control Simulations on a Morphing Wing, Journal of Aircraft, 45:17941803.

Morphing Aircraft
Results), In: Proceedings of AMSE 2010 Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS2010), 28 September1 October, Philadelphia, PASMASIS 2010-3824 (10 pp). Rusnell, M.T., Gano, S.E., Perez, V.M., Renaud, J.E. and Batill, S.M. 2004. Morphing UAV Pareto Curve Shift for Enhanced Performance, 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics & Materials Conference, 1922 April 2004, Palm Springs, CA, AIAA 2004-1682. Sachs, G. and Mehlhorn, R. 1991. Endurance Increase by periodic Optimal Camber Control, In: Proceedings of AIAA Guidance, Navigation and Control Conference, August, New Orleans, LA, AIAA 1991-2670, pp. 635641. Sachs, G. and Mehlhorn, R. 1993. Enhancement of Endurance Performance by Periodic Optimal Camber Control, Journal of Guidance, Control, and Dynamics, 16:572578. Sahoo, D. and Cesnik, C.E.S. 2002. Roll Maneuver Control for UCAV Wing Using Anisotropic Piezoelectric Actuators, In: Proceedings of 43rd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 2225 April, Denver, CO, AIAA 2002-1720 (11 pp). Samareh, J.A., Chwalowski, P., Horta, L.G., Piatak, D.J. and McGowan, A.M.R. 2007. Integrated Aerodynamic/Structural/ Dynamic Analyses of Aircraft with Large Shape Changes, In: Proceedings of 48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2326 April, Honolulu, HI, AIAA 2007-2346. Sanders, B., Cowan, D. and Scherer, L. 2004. Aerodynamic Performance of the Smart Wing Control Effectors, Journal of Intelligent Material Systems and Structures, 15:293303. Santangelo, M., Iuspa, L. and Blasi, L. 2006. Analisi Cineto-Statica di una Struttura Alare Adattiva Hingeless, In: Proceedings of XXXV Convegno Nazionale Associazione Italiana per lAnalisi delle Sollecitazioni (AIAS), 1316 September, Polytechnic University of Marche, Ancona (in Italian) (7 pp). Santhanakrishnan, A., Pern, N.J., Ramakumar, K., Simpson, A. and Jacob, J.D. 2005. Enabling Flow Control Technology for Low Speed UAVs, In: Proceedings of Infotech@Aerospace, 2629 September, Arlington, VA, AIAA 2005-6960 (13 pp). Scott, M.J., Jacob, J.D., Smith, S.W., Asheghian, L.T. and Kudva, J.N. 2009. Development of a Novel Low Stored Volume HighAltitude Wing Design, In: Proceedings of 50th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 45 May, Palm Springs, CAAIAA 20092146 (12 pp). Seigler, T.M., Bae, J.S. and Inman, D.J. 2004. Flight Control of a Variable Span Cruise Missile, In: Proceedings of 2004 ASME International Mechanical Engineering Congress and Exposition, 1319 November, Anaheim, CA, Paper No. IMECE2004-61961. Seigler, T.M., Neal, D.A., Bae, J.S. and Inman, D.J. 2007. Modeling and Flight Control of Large-Scale Morphing Aircraft, AIAA Journal of Aircraft, 44:10771087. Seow, A.K., Liu, Y. and Yeo, W.K. 2008. Shape Memory Alloy as Actuator to Deflect a Wing Flap, In: Proceedings of 49th AIAA/ ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 710 April, Schaumburg, IL, AIAA 2008-1704. Shelton, A., Tomar, A., Prasad, J., Smith, M.J. and Komerath, N. 2006. Active Multiple Winglets for Improved UnmannedAerial-Vehicle Performance, Journal of Aircraft, 43:110116. Shili, L., Wenjie, G. and Shujun, L. 2008. Optimal Design of Compliant Trailing Edge for Shape Changing, Chinese Journal of Aeronautics, 21:187192. Shin, S.J. 2001. Integral Twist Actuation of Helicopter Rotor Blades for Vibration Reduction, PhD Thesis, Massachusetts Institute of Technology. Shin, S.J., Cesnik, C.E.S. and Hall, S.R. 2002. Control of Integral Twist-Actuated Helicopter Blades for vibration Reduction, In: Proceedings of 58th American Helicopter Society Annual Forum, 1113 June, Montreal, Canada, Vol. 2, pp. 21222134.

875

Shkarayev, S., Null, W. and Wagner, M. 2004. Development of Micro Air Vehicle Technology with In-Flight Adaptive Wing Structure, Contractor Report, NASA Langley Research Center, Hampton, VA. Shortal, J.A. and Maggin, B. 1946. Effect of Sweepback and Aspect Ratio on Longitudinal Stability Characteristics of wings at Low Speeds, Langley Memorial Aeronautical Laboratory, NACA TN-1093. Shyy, W., Aono, H., Chimakurthi, S.K., Trizila, P., Kang, C.K., Cesnik, C.E.S. and Liu, H. 2010. Recent Progress in Flapping Wing Aerodynamics and Aeroelasticity, Progress in Aerospace Sciences, 46:284327. Shyy, W., Jenkins, D. and Smith, R. 1997. Study of Adaptive Shape Airfoils at Low Reynolds Number in Oscillatory Flows, AIAA Journal, 35:15451548. Siclari, M.J., Van Nostrand, W. and Austin, F. 1996. The Design of Transonic Airfoil Sections for an Adaptive Wing Concept Using a Stochastic Optimization Method, In: Proceedings of 34th Aerospace Sciences Meeting and Exhibit, 1518 January, Reno, NV, AIAA 1996-0329. Simpson, J., Anguita-Delgado, L., Kawieski, G., Nilsson, B., Vaccaro, V. and Kawiecki, G. 2005. Review of European Research Project Active Aeroelastic Aircraft Structures (3AS), In: Proceedings of European Conference for Aerospace Sciences (EUCASS), 47 July, Moscow, Russia (9 pp). Singh, K. and Chopra, I. 2002. Design of an Improved Shape Memory Alloy Actuator for Rotor Blade Tracking, In: Proceedings of 2002 Society of Photo-Optical Instrumentation Engineers North American Symposium on Smart Structures and Materials, 1721 March, Newport Beach, CA, Vol. 4701, pp. 244266. Singh, V., Skiles, S.M., Krager, J.E., Wood, K.L., Jensen, D. and Sierakowski, R. 2009. Innovations in Design Through Transformation: A Fundamental Study of Transformation Principles, Journal of Mechanical Design, 131: doi: 10.1115/ 1.3125205. Skillen, M.D. and Crossley, W.A. 2006. Developing Morphing Wing Weight Predictors with Emphasis on the Actuating Mechanism, In: Proceedings of 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 14 May, Newport, RI, AIAA 2006-2042 (17 pp). Skillen, M.D. and Crossley, W.A. 2008a. Modeling and Optimization for Morphing Wing Concept Generation II Part I: Morphing Wing Modeling and Structural Sizing Techniques, NASA Report, NASA/CR-2008-214902. Skillen M.D. and Crossley, W.A. 2008b. Morphing Wing Weight Predictors and Their Application in a Template-Based Morphing Aircraft Sizing Environment II, Part II: Morphing Aircraft Sizing Via Multi-Level Optimization, NASA Report, CR-2008- 214903. Smith, K., Butt, J., Spakovsky, M.R. and Moorhouse, D. 2007. A Study of the Benefits of Using Morphing Wing Technology in Fighter Aircraft Systems, In: Proceedings of 39th AIAA Thermophysics Conference, 2528 June, Miami, FL, AIAA 2007-4616 (12 pp). Smith, D.D., Isikveren, A.T., Ajaj, R.M. and Firswell, M.I. 2010. Multidisciplinary Design Optimization of an Active Nonplanar Polymorphing Wing, In: Proceedings of 27th International Congress of the Aeronautical Sciences (ICAS 2010), 1924 September, Nice, France. Smith, J.W., Lock, W.P. and Payne, G.A. 1992. Variable Camber Systems Integration and Operational Performance of the AFTI/ F-111 Mission Adaptive Wing, Technical Memorandum, NASA Dryden Flight Research Center, Edwards, CA. Smith, S.B. and Nelson, D.W. 1990. Determination of the Aerodynamic Characteristics of the Mission Adaptive Wing, Journal of Aircraft, 27:950958. Sofla, A.Y.N., Elzey, D.M. and Wadley, H.N.G. 2008. Two-Way Antagonistic Shape Actuation Based on the One-Way Shape Memory Effect, Journal of Intelligent Material Systems and Structures, 19:10171027.

876

S. BARBARINO ET AL.
Vehicle Design, PhD Dissertation, Aerospace and Mechanical Engineering Department, University of Notre Dame, South Bend, IN. Turnock, S.R. and Keane, A.J. 2009. Morphing of Flying Shapes for Autonomous Underwater and Aerial Vehicles, In: Proceedings of RTO Applied Vehicle Technology Panel (AVT) Symposium, 2024 April, Evora, Portugal (19 pp). Ursache, N.M., Melin, T., Isikveren, A.T. and Friswell, M.I. 2007. Morphing Winglets for Aircraft Multi-Phase Improvement, In: Proceedings of 7th AIAA Aviation Technology, Integration and Operations Conference (ATIO), 1820 September, Belfast, Northern Ireland, AIAA 2007-7813 (12 pp). Ursache, N.M., Melin, T., Isikveren, A.T. and Friswell, M.I. 2008. Technology Integration for Active Poly-Morphing Winglets Development, In: Proceedings of ASME Conference on Smart Materials, Adaptive Structures and Intelligent Systems (SMASIS08), 2830 October, Ellicott City, MD, SMASIS 2008-496 (8 pp). Vale, J., Lau, F., Suleman, A. and Gamboa, P. 2006. Multidisciplinary Design Optimization of a Morphing Wing for an Experimental UAV, In: Proceedings of 11th AIAA/ ISSMO Multidisciplinary Analysis and Optimization Conference, 68 September, Portsmouth, VA, AIAA 2006-7131. Vale, J., Lau, F., Suleman, A. and Gamboa, P. 2007. Optimization of a Morphing Wing Based on Coupled Aerodynamic and Structural Constraints, In: Proceedings of 3rd AIAA MultiDisciplinary Design Optimization Specialists Conference, 2326 April, Honolulu, HI, AIAA 2007-1890 (27 pp). van Dam, C.P. 2002. The Aerodynamic Design of Multi-Element High-Lift Systems for Transport Airplanes, Progress in Aerospace Sciences, 38:101144. Vos, R., Barrett, R., de Breuker, R. and Tiso, P. 2007a. Post-Buckled Precompressed Elements: a New Class of Control Actuators for Morphing Wing UAVs, Smart Materials and Structures, 16:919926. Vos, R., Barrett, R. and Zehr, D. 2008a. Magnification of Work Output in PBP Class Actuators Using Buckling/Converse Buckling Techniques, In: Proceedings of 49th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 711 April, Schaumburg, IL, AIAA 2008-1705. Vos, R., De Breuker, R., Barrett, R. and Tiso, P. 2007b. Morphing Wing Flight Control via Postbuckled Precompressed Piezoelectric Actuators, Journal of Aircraft, 44:10601068. Vos, R., Gurdal, Z. and Abdalla, M. 2008b. A Novel Mechanism for Active Wing Warping, In: Proceedings of 49th AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 711 April, Schaumburg, IL, AIAA 2008-1879. Vos, R., Gurdal, Z. and Abdalla, M. 2010. Mechanism for WarpControlled Twist of a Morphing Wing, Journal of Aircraft, 47:450457. Wang, D.P., Bartley-Cho, J.D., Martin, C.A. and Hallam, B.J. 2001. Development of High-Rate, Large Deflection, Hingeless Trailing Edge Control Surface for the Smart Wing Wind Tunnel Model, In: Proceedings of SPIE Smart Structures and Materials 2001: Industrial and Commercial Applications of Smart Structures Technologies, 58 March, Newport Beach, CA, Vol. 4332, pp. 407418. Wang, J.M., Jones, C.T. and Nixon, M.W. 1999. A Variable Diameter Short Haul Civil Tiltrotor, In: Proceedings of 55th Annual Forum of the American Helicopter Society, 2527 May, Montreal, Canada (8 pp). Warwick, G. 2009. Morphing Methods, Aviation Week & Space Technology, p. 55. Waszak, M.R., Jenkins, L.N. and Ifju, P.G. 2001. Stability and Control Properties of an Aeroelastic Fixed Wing Micro Aerial Vehicle, In: AIAA Atmospheric Flight Mechanics Conference, 69 August, Montreal, Canada, AIAA 2001-4005 (11 pp). Weiss, P. 2003. Wings of Change: Shape-Shifting Aircraft may Ply Future Skyways, Science News, 164:359, 362 and 365.

Sofla, A.Y.N., Meguid, S.A., Tan, K.T. and Yeo, W.K. 2010. Shape Morphing of Aircraft Wing: Status and Challenges, Materials and Design, 31:12841292. Song, G. and Ma, N. 2007. Robust Control of a Shape Memory Alloy Wire Actuated Flap, Smart Materials and Structures, 16:N51N57. Song, A., Tian, X., Israeli, E., Galvao, R., Bishop, K., Swartz, S. and Breuer, K. 2008. Aeromechanics of Membrane Wings with Implications for Animal Flight, AIAA Journal, 46:20962106. Song, D., Wang, H., Zeng, L. and Yin, C. 2001. Measuring the Camber Deformation of a Dragonfly Wing Using Projected Comb Fringe, Review of Scientific Instruments, 72:24502454. Spangler, R.L. and Hall, S.R. 1990. Piezoelectric actuators for helicopter rotor control, In: Proceedings of 31st AIAA/ASME/ ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 24 April, Long Beach, CA, AIAA1990-1076. Spillman, J. 1992. The Use of Variable Camber to Reduce Drag, Weight and Costs of Transport Aircraft, Aeronautical Journal, 96:19. Stanewsky, E. 2001. Adaptive Wing and Flow Control Technology, Progress in Aerospace Sciences, 37:583667. Stanford, B., Abdulrahim, M., Lind, R. and Ifju, P. 2007. Investigation of Membrane Actuation for Roll Control of a Micro Air Vehicle, Journal of Aircraft, 44:741749. Steadman, D.L., Griffin, S.L. and Hanagud, S.V. 1994. StructureControl Interaction and the Design of Piezoceramic Actuated Adaptive Airfoils, In: Proceedings of AIAA/ASME Adaptive Structures Forum, 2122 April, Hilton Head, SC, AIAA 19941747, pp. 135142. Straub, F., Anand, V.R., Birchette, T. and Lau, B.H. 2009. SMART Rotor Development and Wind Tunnel Test, In: Proceedings of 35th European Rotorcraft Forum 2009 (ERF 2009), 2225 September, Hamburg, Germany, Article 1200 (21 pp). Straub, F.K., Ngo, H.T., Anand, V. and Domzalski, D.B. 2001. Development of a Piezoelectric Actuator for Trailing-Edge Flap Control for Full Scale Rotor System, Smart Materials and Structures, 10:2534. Strelec, J.K., Lagoudas, D.C., Khan, M.A. and Yen, J. 2003. Design and Implementation of a Shape Memory Alloy Actuated Reconfigurable Airfoil, Journal of Intelligent Material Systems and Structures, 14:257273. Studebaker, K. and Matuska, D. 1993. Variable Diameter Tiltrotor Wind Tunnel Test Results, In: Proceedings of American Helicopter Society 49th Annual Forum, May, St. Louis, MO. Sullivan, J. and Watkins, W. 2003. Obliquing Wing with Extendable Span Critical Design Review, Critical Design Review, AAE 490T/590T AT 490D Tech 581N Design/Build/Test, Purdue University, IN. Supekar, A.H. 2007. Design, Analysis and Development of a Morphable Wing Structure for Unmanned Aerial Vehicle Performance Augmentation, Masters Thesis, University of Texas at Arlington, TX. Szodruch, J. 1985. The Influence of Camber Variation on the Aerodynamics of Civil Transport Aircraft, In: Proceedings of 23rd Aerospace Sciences Meeting, 1417 January, Reno, NV, AIAA 85-0353. Szodruch, J. and Hilbig, R. 1988. Variable Wing Camber for Transport Aircraft, Progress in Aerospace Sciences, 25:297328. Thill, C., Etches, J., Bond, I., Potter, K. and Weaver, P. 2008. Morphing Skins, The Aeronautical Journal, 112:117139. Thill, C., Etches, J.A., Bond, I.P., Potter, K.D. and Weaver, P.M. 2010. Composites Corrugated Structures for Morphing Wing Skin Applications, Smart Materials and Structures, 19:124009. Thornton, S.V. 1993. Reduction of Structural Loads Using Maneuver Load Control on the Advanced Fighter Technology Integration (AFTI)/F-111 Mission Adaptive Wing, NASA TM 4526. Torres, G.E., 2002. Aerodynamics of Low Aspect Ratio Wings at Low Reynolds Numbers with Applications to Micro Air

Morphing Aircraft
Weisshaar, T.A. 2006. Morphing Aircraft Technology New Shapes for Aircraft Design, RTO-MP-AVT-141, Neuilly-sur-Seine, France. Wereley, N.M. and Gandhi, F. 2010. Flexible Skins for Morphing Aircraft, Journal of Intelligent Material Systems and Structures, 21:16971698. Wickramasinghe, V.K., Chen, Y., Martinez, M., Kernaghan, R. and Wong, F. 2009. Design and Verification of a Smart Wing for an Extremely-Agile Micro-Air-Vehicle, In: Proceedings of 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2132 (11 pp). Wiggins, L.D., Stubbs, M.D., Johnston, C.O., Robertshaw, H.H., Reinholtz, C.F. and Inman, D.J. 2004. A Design and Analysis of a Morphing Hyper-Elliptic Cambered Span (HECS) Wing, In: Proceedings of 45th AIAA/ASME/ASCE/ AHS/ASC Structures, Structural Dynamics and Materials Conference, 1922 April, Palm Springs, CA, AIAA 2004-1885 (10 pp). Wilbur, M.L. and Wilkie, W.K. 2004. Active-Twist Rotor Control Applications for UAVs, Report, US Army Research Laboratory Vehicle Technology Directorate, Hampton, VA. Wildschek, A., Gru newald, M., Maier, R., Steigenberger, J., Judas, M., Deligiannidis, N. and Aversa, N. 2008. Multi-Functional Morphing Trailing Edge Device for Control of All-Composite, All-Electric Flying Wing Aircraft, In: Proceedings of The 26th Congress of International Council of the Aeronautical Sciences (ICAS), 1419 September, Anchorage, Alaska, AIAA 20088956.

877

Wilson, J.R. 2002. Active Aeroelastic Wing: A New/Old Twist on Flight, Aerospace America, 40:3437. Wittmann, J., Steiner, H.-J. and Sizmann, A. 2009. Framework for Quantitative Morphing Assessment on Aircraft System Level, In: Proceedings of 0th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA J. 2009-2129 (18 pp). Yang, S.M., Han, J.H. and Lee, I. 2006. Characteristics of Smart Composite Wing with SMA Actuators and Optical Fiber Sensors, International Journal of Applied Electromagnetics and Mechanics, 23:177186. Yu, Y., Li, X., Zhang, W. and Leng, J. 2007. Investigation on Adaptive Wing Structure based on Shape Memory Polymer Composite Hinge, In: Proceedings of SPIE International Conference on Smart Materials and Nanotechnology in Engineering, 1 July, Harbin, China, Vol. 6423. doi:10.1117/ 12.779394 (7 pp). Zhang, Q. and Hoffmann, F. 2009. Benefit Studies for Rotor with Active Twist Control Using Weak Fluid-Structure Coupling, In: Proceedings of 35th European Rotorcraft Forum, 2225 September, Hamburg, Germany, Article 1217. Zhao, L., Dawes, W.N., Parks, G., Jarrett, J.P. and Yang, S. 2009. Robust Airfoil Design with Respect to Boundary Layer Transition, In: Proceedings of 50th AIAA/ASME/ASCE/AHS/ ASC Structures, Structural Dynamics, and Materials Conference, 47 May, Palm Springs, CA, AIAA 2009-2273 (11 pp).

You might also like