You are on page 1of 9

Proceedings of the 10th Mediterranean Conference on Control and Automation - MED2002 Lisbon, Portugal, July 9-12, 2002.

AERODYNAMICALLY COUPLED FORMATION FLIGHT OF AIRCRAFT


D. F. Chichka1,2 , J. D. Wolfe1,2 , and J. L. Speyer1,3
University of California, Los Angeles, Los Angeles, California, USA chichka@ucla.edu, wolfe@talus.seas.ucla.edu, speyer@seas.ucla.edu
1

e-mail:

Keywords: Formation ight, Aerodynamic cou- 1 Introduction pling, Coordinated control. Close formation ight of aircraft can be of benet in many situations. It has long been known[1], Abstract for example, that by ying in a proper formation, We examine the ight of aircraft in close forma- aircraft can achieve aerodynamic eciency much tion, such that the aerodynamic eects of a lead greater than is possible when ying alone. Such aircraft on a follower are pronounced. These ef- an eect might be of interest to commercial avifects may include a reduction in drag, and un- ation to reduce fuel costs on long ights to comdesireable, strong rolling moments produced on mon destinations, or to the military in an eort to trailing aircraft in such a formation. These eects extend the range of airborne sorties. Such appliare nonlinear with regard to the relative positions cations may involve both manned and unmanned of the aircraft. In the case of a formation ight aerial vehicles (UAVs). for drag reduction, the result is an unstable ight condition for the trailing craft. An additional difculty is that changes in the wake of a leader due to its motion take some time to propagate to the trailing craft. This introduces a time delay in the system that complicates controller synthesis. Furthermore, as the number of aircraft increases, the control problem may also become more complex due to communication bandwidth or other constraints, so that a centralized solution is dicult to implement. This paper considers formations of three aircraft. Controllers are derived using linear system models, and compared in a simulation that includes both the aerodynamic interactions and the delay. It is seen that the delay causes signicant complications in controller design. Other applications may include close formation ight only as a part of the mission. Aerial refueling, for instance, involves relatively small aircraft ying well within the aerodynamic wash of a tanker craft. The coordinated ight of unmanned aerial combat vehicles (UCAVs) with each other or in support of manned aircraft, may include manuevers that take the vehicles into the wakes of other vehicles. For all these reasons, it is important to understand the control of aircraft in such aerodynamically coupled situations. These areas have been addressed by a number of authors in recent years. Formation ight for drag reduction has been examined both analytically and experimentally for two aircraft by Hummel[2], who reported a reduction in drag of at least fteen per cent. The use of peak-seeking control to maximize the eciency gain for a pair of aircraft is investigated in [3], and a more general examination of controllers for such cases is given in [4]. In all of these works, the analysis is conned to

Assistant Research Engineer, Mechanical and Aerospace Engineering. 3 Professor, Mechanical and Aerospace Engineering.

the case of two aircraft, in which one is a leader and the other takes position to the rear and o to the side of the leader. An assumption common to each paper is that the lead craft maintains straight and level ight, and is unaected by the trailing aircraft. The problem therefore reduces to controlling the follower craft in the presence of the wake of the leader. This assumption is not tenable in formations of more than two vehicles, such as are posited in [5], where large formations of high-altitude solar-electric craft are proposed for near-innite duration ight. Formations of three to ve such vehicles are addressed in [6], where the use of decentralized controllers for such formations is addressed.

2.1

Horseshoe Vortex Approximation.

The horseshoe vortex approximation assumes that the aerodynamic presence of the craft is primarily due to the wing. The wing is modeled as a bound vortex, and the wake is represented by a pair of vortices, one trailing from near each wingtip, as sketched in Figure 1. To fulll the requirements of Helmholz vortex laws, these trailing vortices extend to innity.

P The present paper examines the problem of deh riving controllers for formation of more than two vehicles. The specic problem addressed is the creation and maintenance of a specic formation in which drag is reduced or minimized. Earlier work is extended by including a more realistic model of the aircraft wakes. Among other eects, this introduces a delay between the time at which a lead aircraft moves and when its eect is felt on the aerodyamics of a following vehicle. The vehicles simulated are representative of light private Figure 1: Horseshoe vortex description of formaairplanes. tion ight. The next section of the paper includes a brief overview of formation ight dynamics, and de- Let P be some point on the wing of the trailing scribes the model used in these investigations. aircraft. The eect of a trailing vortex at P is Section 3 discusses the derivation of controllers given by the Biot-Savart law to be for the problem. Simulation results are presented in Section 4, and the paper concludes with a nal q= (1 cos ) (1) 4h section.

Aerodyamic Modeling

The aerodynamics of formation ight have been covered in many books and papers (see for example [1, 2, 5]), so only a very brief discussion is included here. We begin with the horseshoe vortex model of Prandtl[7], which is both straightforward and allows a discussion of all major eects.

where h is the perpendicular distance to the vortex, which is assumed straight. The induced velocity q is normal to the plane dened by P and the vortex, and is positive by the right hand rule about the vortex. These directions are shown by the arrows on the trailing vortices in the gure; it is seen that the vortex from the right wingtip of the lead plane will induce a velocity out of the page on the trailing craft. Because the eects fall o as 1/h, it is clear that most of the formation eect is due to the nearest trailing vortex. From the geometry, it is easily seen that the dominant

eects will be an overall upwash on the trailing craft. The upwash is strongest near the inboard wigtip, leading to a strong outward rolling moment. Because the upwash leads to a reduction in drag, this reduction also is strongest near the inboard wingtip, so that the trailing craft also experiences a slight outward yawing moment. For the purposes of this paper, it is assumed that the other induced forces and moments, such a pitching moment and side force, are negligible. All of the trailing vortices act on all of the lifting surfaces of all craft, as do the bound vortices that stand in for the wings. However, if the trailing craft is appreciably behind the leader, its eects on the leader are slight, and usually considered negligible.

wing. This is the model used in the present research. The wing is modeled using vortex ring panels along the camber line, and the trailing vortex strengths are determined from the circulation of the trailing edge panels. For an excellent description of such methods, see Katz and Plotkin[8]. To better describe the vorticity distribution near the wingtips, cosine spacing (as opposed to equal sized panels) is used laterally, as suggested by Moran[9]. Wing thickness is not modeled. 2.3 Wake Propagation

A usual assumption of the horseshoe vortex model is that the trailing vortices are straight and extend to innity. In fact, the vortex wake from an aircraft is a function of where the craft 2.2 Distributed Wake Approximation was when the wake was generated. The wake also moves due to self-induced velocity and motion inThe horseshoe vortex description allows a quick understanding of the formation eects. From a duced by wind or other sources. distance, it is also a fair description of the aircraft In much of the work cited above, only two aircraft wake, particularly once the wake has had time to were considered, and the intent was to put a folroll up. Near the trailing edge of the wing, lower into position relative to the leader. With however, the wake is better described by a set of the assumption that the leader ies a straight vortices of varying strengths. and level path, and ignoring any ambient wind, the only eects on the wake position will be selfinduced velocity and any eects of the follower on the wake of the leader. The self-induced motion is fairly well-understood, and can in any case be considered a constant unknown bias. The eect of the trailing aircraft is a function of its behavior. Some preliminary work suggests that this eect is likely to be small. In formations of more than a pair of aircraft, or in which the aircraft work together, the assumption of straight and level ight for the leader is no longer valid. Instead, the eect of motion of a leader on a follower must be considered. Unfortunately, this is not straightforward, because the eect of the leader is due to its wake, rather than its immediate position. For simplicity, we here assume that the wake reThis is shown in Figure 2, where for clarity the mains where it is created; that is, it follows pretrailing vortices are slightly separated from the cisely the path of the vehicle that created it. A modied horseshoe approximation of the wake Figure 2: Two aircraft with distributed wakes.

angles with respect to a local horizontal frame (, , ), and a position in inertial space (x, y, z). A linear model of the formation dynamics about a trim condition may be constructed by calculating central dierences of the simulation dynamics derivative routine. The resulting linear model may then be used to construct linear controllers. There is a signicant eect that is ignored by this approach the dynamics of the wake are not modelled. When central dierences about the nominal trajectory are taken, the wakes of the planes do not move. This causes the resulting linear system to exhibit dependencies on the absolute states of the planes. In reality, these terms of the linear system are dependent on the position of the planes relative to the wakes of the other planes, not the planes position in free space. To correctly account for this requires additional dynamics either the wake dynamics must be modelled, or wake location can be taken as the delayed position of leading planes (in which case a Pad e approximation of the delay may be used). To avoid the additional modelling, we simply allow the dynamics to have relationships to absolute position, with the understanding that these eects assume a perfectly steady nominal trajectory. 3.2 Linear Controllers

Figure 3: Trailing wakes with arbitrary aircraft motion. therefore may look something like Figure 3. The eect of such a wake from the lead craft on the follower will be much dierent than would the straight wakes shown in Figure 1. In fact, the dominant eects of the wake are due to the small portion of the wake nearest the trailing aircraft. This introduces a delay in the eects of leader motion on the dynamics of the trailing craft. The amount of the delay is governed by the speed of the formation and the longitudinal separation of the individual aircraft. For vehicles at high speeds with close following distances, it might be all but negligible. For lower-speed craft with several wingspans of separation, it may be measured in seconds.

We create control gains for the aircraft assuming that the states of the planes are known perfectly. The design of state estimators for formation ight is a complicated issue that has been considered in other papers [10, 11]. Sensor systems based on these ideas have been experimentally veried to 3 Controller Synthesis estimate relative position to within several cenIn this section, we introduce some considerations timeters. of controller synthesis for the formation ight The simplest approach for controller synthesis is problem, and discuss the controllers derived for to implement a linear quadratic regulator (LQR) this study. upon the combined state spaces of all the planes present in the formation. As a result, each control surface of each plane will deect based on Each craft in the formation is described by 12 information about every state of every plane in states, which comprise a velocity vector (u, v, w), the formation. an angular velocity vector (p, q, r), three Euler 3.1 Linear Model

Because the information ow required for such a centralized controller may be prohibitive, we have considered several decentralized approaches to controller synthesis. A fully decentralized controller may be designed by generating a LQR for each craft from the portion of the dynamics matrices corresponding only to the 12 states of that plane. All eects from the other vehicles are then regarded as disturbances to the craft. Alternatively, xed structure control design procedures [12] may be used to create a nearest neighbor controller, where each craft has gains on its own states and those of the craft that is immediately ahead.

less signicant than the motion of the lead aircraft for relatively small following distances. 4.2 The Linear System

To generate controllers for this study, the formation was rst trimmed at the nominal condition. A linear system was created using central dierences. Recall that due to the time delay, small perturbations of the lead craft have almost zero immediate eect on a trailing craft. Controllers were derived using Scilab2 .
imag. axis 7 5 3

transmission zeros and poles

4
4.1

Simulation Results
A Light Personal Aircraft

The aircraft modeled for this study is based upon aircraft A from Roskam[13]. The vehicle is representative of a small four-place personal transportation airplane1 . The wingspan is 35.8 feet, and the vehicle weighs 2645 lbs. The ight speed is 220 feet per second. A following distance of 36 feet, or just over one wingspan, creates a delay of about 0.16 second in the eects of a lead aircraft on the follower. While small, this may still be signicant given the dynamics of the airplane. Following distances of several wingspans, likely in practice, will create much longer delays.

-1 -3 -5 -7 -19

real axis 1 5

-15 Poles

-11

-7

-3

Figure 4: Poles for three aircraft without coupling. It is worthwhile to consider the formation eects in the linear system. Figure 4 displays the poles of the open loop dynamical system for three aircraft without coupling. Because the coupling is absent, the system poles for each aircraft lie directly on top of each other, and the gure appears to display only a single aircraft. Note the two slightly unstable poles on the real axis.

For the purposes of this study, only the lifting and control surfaces of the aircraft are modeled. All surfaces are modeled using vortex panels to simulate the camber line; no thickness is included. Cosine spacing is used on the wings, so that the vorticity distribution near the wingtips may be more accurately modeled. The tail surfaces are When the aerodynamic coupling is added, the remodeled with evenly distributed rings. This al- sult is Figure 5. Note that because the coupling is one-way, from leader to follower, the poles for the lows a better modeling of the control surfaces. rst craft in the formation do not move. Several The trailing wakes from the wings are modeled as of the poles for the other craft shift only slightly. stationary where created. This captures the deSome, however, jump dramatically, leading to the lays inherent in formation ight eects, without additional unstable modes seen near the unit cirattempting to completely model wake roll-up or 2 self-induced motion. These are considered to be Scilab is a free tool, similar in form and ca1

Roskam[13], p.590.

pability to Matlab (tm). It is available from http://www-rocq.inria.fr/scilab/.

transmission zeros and poles imag. axis 7 5

1 -1

tion of the second vehicle. The desired position is 36 feet of longitudinal separation and 36 feet lateral, such that the craft are ying almost wingtips aligned. This is very close to being the optimal lateral position for minimum drag, and thus provides a logical test case. The initial condition of the simulation is an overlap of ten feet for the second craft in the formation. In this position, the craft is well into the downwash eld behind the rst craft. The outward moment induced is much reduced from the nominal, and the lift induced is also much lower.

-3

-5

real axis

-7 -19 Poles -15 -11 -7 -3 1 5

Figure 5: Poles for three aircraft with aerodynamic coupling.


second plane lateral position (ft)

38 36 34 32 30 28 26 0 2 4 6 8 10 time (sec) 12 14 16
centralized decentralized (coupling effects considered) decentralized (coupling effects ignored) nearest neighbor expected linear response

cle. These modes appear to be due to the nature of the induced forces on the trailing craft. Because the induced lift, drag, and moments are strongly aected by lateral position, they create oscillation. An uncontrolled follower will soon begin to wobble, and before long will roll away from the formation. 4.3 Results

Figure 7: Lateral position of second of three aircraft. Figure 7 shows the lateral position of the second of the three aircraft using various controllers. The solid line represents a controller derived using the linear system derived at the nominal condition, with full state knowledge available. Two decentralized controllers are represented, one derived from a linearization about the nominal formation (coupling eects considered), and one from linearization of the craft ying singly (coupling effects ignored). The nearest neighbor controller is a leader-follower; the aircraft looks only at its immediate leader. The expected linear response is that predicted by the linear simulation used to create the controllers.

Figure 6: Nominal and perturbed formation.

This section presents results of simulations run Compared to the expected linear response, only using several dierent controllers. The nominal the controller without coupling considered shows positions of the aircraft are shown in Figure 6, similar capability. The rise time is nearly as where the dashed outline shows the initial condi-

good, though the aircraft then oscillates much more than expected. The oscillations die out at a bit over ten seconds, which is also the time at which the other controllers achieve nominal position. All of the other controllers have almost identical response characteristics, beginning with a fairly rapid initial response and proceeding with a more gradual rise to the nominal, without overshoot. Each controller nally ends with a slight bias.

72.3 72.1

third plane lateral position (ft)

71.9 71.7 71.5 71.3 71.1 70.9 70.7 70.5 0 2 4 6


centralized decentralized (coupling effects considered) decentralized (coupling effects ignored) nearest neighbor expected linear response

0.25

8 10 time (sec)

12

14

16

antisymmetric aileron deflection (deg)

0.00 -0.25 -0.50 -0.75 -1.00 -1.25 -1.50 -1.75 -2.00

centralized decentralized (coupling effects considered) decentralized (coupling effects ignored) nearest neighbor

Figure 9: Lateral position of third of three aircraft.

8 10 time (sec)

12

14

16

Figure 8: Anti-symmetric aileron deection of second aircraft. The control behavior is reected in the antisymmetric aileron activity, shown in Figure 8. The controller without coupling eects drives the ailerons hard, while the other controllers have a similar initial deection to begin the recovery. After the initial peak, however, these controllers relax and allow the wake-induced moment to help roll the craft back out to the desired position, while the rst oscillates the ailerons to make up for the unexpected wake eects. One reason that the recovery for the other controllers might be so slow is that the linear system on which they are based shows a coupling of moment with lateral position. From the linear system, it would be expected that the outward rolling moment increases with overlap, which is not the case in actuality. In eect, the controller has added damping to the system to account for the moment dependency on position, and the result is an overdamped system when the moment is not seen.

It is also interesting to note the eect of the second aircrafts motion on the third vehicle in the formation. Initially, as shown in Figure 9, the third airplane rolls inward, due to the lack of the expected moment from the second. The controller without coupling eects considered reponds most quickly, but then oscillates around the nominal position. The oscillations die out only slowly, being forced by the wake eects from the second craft. Note that the expected linear response is almost zero, because the eect from the second plane is not seen in the linear system. Again, all the other controllers respond almost identically. They again show a delay in responding, followed by a highly damped rise to a stable position. Again, this appears to be due to the damping induced by the controller. With the second craft out of position, the expected rolling moment is not seen by the third. Note that as the second plane moves back into position, recreating the expected forces, the third vehicle responds more quickly. At 6 seconds, the third craft is seeing a wake created at 34 feet, while it itself is over a foot inward. Thus the aerodynamic system is that of only one foot error, the moment is nearly restored, and the system is no longer so severely overdamped.

Conclusions

In addressing the problem of aerodynamically coupled formation ight, we have included the delay in the eects of a lead aircraft on a follower. This causes some diculty in create a linear system for control design, in that small perturbations in a lead aircraft position have no immediate eect on the dynamics of a trailing craft. However, the eect is strong after a delay. One eect of this is that while the natural choice of state space for controller design would include relative positions between the craft, this choice is not tenable. The aerodynamic interactions are functions of relative position only in steady state. Controllers have been derived for this study that do not include the delay. They work acceptably, in that they stabilize the system and eventually reject errors. However, the response is usually slow, and if not slow is unacceptably oscillatory. This indicates that the aerodynamic coupling should be better included in the controller design. However, this will require inclusion of the delay in the synthesis. This is the subject of ongoing investigation.

and Development, 7 Rue Ancelle, 92200 Neuilly-sur-Seine, France, 1996. [3] D.F. Chichka, J.L. Speyer and C.G. Park, Peak-Seeking Control with Application to Formation Flight, Proceedings of the 38th Conference on Decision and Control, Phoenix, AZ, December 1999. [4] Meir Pachter, John J. DAzzo, and Andrew W. Proud, Tight Formation Flight Control, AIAA Journal of Guidance, Control, and Dynamics, Vol. 24, No. 2, March-April 2001, pp. 246-54. [5] D. Chichka and J. Speyer, Solar-Powered Formation-Enhanced Aerial Vehicle Systems for Sustained Endurance, Proceedings of the 1998 American Control Conference, Philadelphia, Pennsylvania, June 1998. [6] J. D. Wolfe, D. F. Chichka, and J. L. Speyer, Decentralized Controllers for Unmanned Aerial Vehicle Formation Flight, presented at the AIAA Guidance, Navigation, and Control Conference, San Diego, California, July, 1996. [7] Prandtl and Tietjens, Applied Hydro- and Aeromechanics, Dover, New York, NY, 1957. (Reprint of 1934 edition.) [8] Joseph Katz and Allen Plotkin, Low-Speed Aerodynamics, 2nd. Ed., Cambridge University Press, Cambridge, UK, 2001. [9] Jack Moran An Introduction to Theoretical and Computational Aerodynamics, John Wiley and Sons, New York, NY, 1984.

Acknowledgments
This work was partially supported by the US Air Force Oce of Scientic Research under grant F49620-01-1-0361, and by NASA Ames Research Center under grant NAG2-1484.

References

[1] E. L. Houghton and P. W. Carpenter, Aerodynamics for Engineering Students, John Wiley and Sons, New York, NY, 4th ed., [10] W. R. Williamson, M. F. Abdel-Hafez, 1993. I. Rhee, E. J. Song, J. Wolfe, D. Cooper, D. Chichka, and J. L. Speyer, An In[2] D. Hummel, The Use of Aircraft Wakes strumentation System Applied to Formato Achieve Power Reductions in Formation tion Flight, in Proceedings of the AIAA 1st Flight, in The Characterisation & ModicaTechnical Conference and Workshop on Untion of Wakes from Lifting Vehicles in Flumanned Aerospace Vehicles, Systems, Techids, AGARD Conference Proceedings 584, nologies, and Operations, Portsmouth, VirThe Advisory Group for Aerospace Research ginia, May 2002.

[11] W. Williamson, T. Rios, and J. L. Speyer, Carrier Phase Dierential GPS/INS Positioning for Formation Flight, in Proceedings of the 1999 American Control Conference, San Diego, California, June 1999, Vol. 5, pp. 3665-70. [12] J. D. Wolfe and D. F. Chichka, An Effective Design Algorithm for Optimal Fixed Structure Control, Proceedings of the 36th IEEE Conference on Decision and Control, San Diego, California, December 1997, Vol 3, pp. 2625-27. [13] Jan Roskam, Aircraft Dynamics and Control, Roskam Aviation and Engineering Corporation, Ottawa, Kansas, 1979, second printing.

You might also like