DYNAMICS OF A MARS *

Håvard Fjær Grip⋆, Wayne Johnson⋆⋆, Carlos Malpica⋆⋆, Daniel P. Scharf⋆, Milan Mandić⋆, Larry Young⋆⋆, Brian Allan†, Bérénice Mettler‡, and Miguel San Martin⋆

⋆Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA ⋆⋆NASA Ames Research Center, Moffet Field, CA †NASA Langley Research Center, Hampton, VA ‡University of Minnesota, Aerospace Engineering and Mechanics, Minneapolis, MN

Abstract

Helicopters have the potential to transform Mars exploration by providing a highly mobile platform for forward reconnais- sance as an aid for ground-based systems. NASA is there- fore considering the possibility of sending a small helicopter to the Martian surface as part of a future mission. Helicopter flight on Mars is challenging due to the extremely thin atmo- sphere, which is only partially offset by a reduction in gravity. In this paper we focus on flight dynamics and controllability issues for the proposed Mars Helicopter, in particular the ar- eas in which the dynamics departs from typical behavior on Earth. We discuss insights gained from modeling and simu- lation as well as system identification performed with a test Figure 1: Artist’s concept of the proposed Mars Helicopter. vehicle in the relevant atmospheric condition, which culmi- nated in the first demonstration of controlled helicopter flight in Martian atmospheric conditions in May 2016. 1.1 Challenges of Mars Helicopter Flight

Like any spacecraft or spacecraft instrument, a helicopter 1 Introduction designed for Mars faces a host of challenging requirements not typically seen on Earth: it must withstand high structural Starting with the first attempted flybys of Mars in the 1960s, loads during launch, extreme temperature variations, high human exploration of the Red Planet has evolved through levels of radiation, and be vacuum compatible; it must sat- ever-more sophisticated means, with the use of orbiters, isfy cleanliness requirements related to planetary protection; stationary landers, and more recently rovers that have trav- and it must operate entirely without physical intervention af- eled over distances of tens of kilometers in search of new ter launch. It must also be compactly designed and safely knowledge. Yet, despite discussion since the early days of deployable after landing. space exploration, no mission has so far attempted to un- In this paper, we focus on a different set of challenges, lock the aerial dimension of Mars exploration through the namely, those related to the flight dynamics of the vehicle use of atmospheric flyers. when operating in the Martian environment, and how these This could be about to change, as NASA is considering affect the mechanical design of the vehicle and the flight the possibility of sending a small helicopter to the Martian control algorithms. Two aspects of the environment are pri- surface as part of a future mission, as a technology demon- mary drivers for the flight dynamics of a helicopter on Mars: stration to verify the feasibility and utility of using • Atmosphere: the Martian atmosphere consists primar- for future Mars exploration. ily of CO at only 1–2% of Earth’s atmospheric den- The use of helicopters promises to bridge a resolution gap 2 sity at sea level, equivalent to altitudes of approxi- in current Mars exploration capabilities—between orbiters mately 100, 000 ft on Earth. providing large-area imagery at low resolution, and rovers that provide detailed imagery limited by line-of-sight from the • Gravity: the Martian gravity is approximately 38% of current rover location. Paired with a rover, a helicopter can Earth’s gravity. act as a forward reconnaissance platform, helping to identify promising science targets or mapping the terrain ahead of The most obvious effect of the reduced density is a reduc- the rover. Looking further ahead, helicopters may one day tion in lift capability for any given rotor. The reduced grav- carry their own science payloads to areas that are inacces- ity, while helpful, does not nearly make up for this reduction sible to rovers. in lift. Beyond this, the Martian environment also influences the helicopter flight dynamics in less obvious but highly con- *The work of H. F. Grip, D. P. Scharf, M. Mandić, and M. San Martin was carried out at the Jet Propulsion Laboratory, California Institute of Technology, sequential ways, which ultimately influence both the design under a contract with the National Aeronautics and Space Administration. of the helicopter itself and the algorithms used to control it.

© 2017. All rights reserved. • sensitivity to edgewise flow

• low-frequency dynamics, including apparent inertia ef- fects due to rotor flapping

• helicopter stability properties

• additional insights from system identification in Martian atmospheric conditions We end with some concluding remarks on remaining work in the flight dynamics area for the proposed Mars Helicopter.

1.4 Previous Work on Mars Rotorcraft Although aerial vehicle concepts for Mars have been pro- Figure 2: In this image, the Mars Helicopter demonstration vehi- posed since the earliest days of space exploration, the full cle is taking off for the first controlled flight in Martian atmospheric challenges of flight in the extremely thin, cold, and primarily conditions, inside JPL’s 25-ft Space Simulator. CO2-based atmosphere of Mars only become apparent after the Viking Lander missions of the 1970s. For approximately Designing a helicopter for Mars also presents serious two decades afterwards only Mars concepts were challenges in terms of testing, verification, and validation. studied. Savu and Trifu [12] were among the first to con- It is not practically possible to fully replicate the Mars envi- sider the design implications of Mars rotorcraft; however, ronment on Earth; this forces a greater reliance on analysis, their work was not carried forward beyond its initial study. modeling, and simulation than for a typical small-scale ro- In 2000, Stanford University and JPL, as a tangential off- torcraft program on Earth, combined with limited testing in shoot of Stanford’s NASA NIAC-sponsored micro-rotorcraft partially replicated environments. work attempted a proof-of-concept test of a small rotor under Mars-like atmospheric conditions in a JPL vacuum cham- 1.2 Controlled-Flight Demonstration ber [7]. No experimental results were published from this initial proof-of-concept test, though video was released at In May 2016, the Mars Helicopter project conducted the first- the time. Independently, at approximately the same time, a ever controlled helicopter flight in Martian atmospheric con- series of papers were published by NASA Ames Research ditions, in the 25-ft Space Simulator at NASA’s Jet Propul- Center studying the technical challenges of various Mars ro- sion Laboratory. The demonstration vehicle was a co-axial torcraft configurations and other vertical lift planetary aerial helicopter featuring a 1.21-m diameter rotor—the same size vehicles [1, 18, 24, 20]. being considered for Mars flight—rotating at a speed of In 2000, Sikorsky and NASA Ames Research 2600 RPM (see Figures 2 and 4). The vehicle was built Center co-sponsored the AHS International Student Design by AeroVironment, Inc., the primary industry partner on the Competition on the topic of a Mars rotorcraft. Thompson Mars Helicopter project. Because the vehicle had to lift its [15] and Datta et al. [3] are summary papers describing own weight on Earth, it did not carry batteries or an avionics respectively the Georgia Tech and University of Maryland computing platform, but was connected to a ground station Mars rotorcraft design entries. In 2002, the first experimen- via power and data tethers hanging below the vehicle. tal results for a rotor hover test under Mars-like conditions The flight demonstration followed an intensive vehicle de- were published [26]. Over the next decade-and-a-half, pa- sign effort that was driven by analysis of the flight properties pers continued to be occasionally published by NASA Ames of helicopters on Mars. Moreover, due to a high degree of [19, 25, 21, 22, 23] on the topic of planetary aerial vehicles, modeling uncertainty and the almost complete lack of prior both fixed- and vertical lift aerial vehicles. work on flight dynamics in Mars-like conditions, a system As time went on, interest slowly grew, and papers by other identification program was executed to identify the actual authors, organizations, and countries began to be published dynamics of the vehicle as-built, thereby casting further light on Mars rotorcraft and vertical lift planetary aerial vehicles on the unique properties of Mars helicopters. [17, 13]. In 2014, JPL and AeroVironment published an ini- tial paper on what would ultimately become the conceptual 1.3 Topics of this Paper foundation for the current Mars Helicopter effort [2]. The concept has significantly evolved over the past few years, In this paper we focus on flight dynamics and controllabil- but retains many of its core features. ity issues for the Mars Helicopter, in particular the areas Although several of the papers cited were concerned with in which the Mars departs from typical feasibility of Mars rotorcraft from the point of view of thrust behavior on Earth. The following areas are covered specif- generation and power consumption, they did not address ically: the flight dynamics or the design of closed-loop control for • overview of the Mars Helicopter vehicle design Mars rotorcraft at a detail level. • modeling of Mars Helicopter flight dynamics 2 Vehicle Overview • impact of low density on flap dynamics in the Martian environment, and consequences for control and vehicle The current Mars Helicopter baseline design features a co- design axial rotor, with two counter-rotating, hingeless, 2-bladed ro-

© 2017. All rights reserved. Safety tether (removed for free-flight)

Visual targets for motion tracking

Counter-rotating rotors

Motor, gears & pitch linkage assembly

Electrical cables Landing legs

Figure 3: CAD drawing of the proposed Mars Helicopter. Figure 4: The demonstration vehicle used for controlled-flight demonstration in Martian atmospheric conditions. tors measuring 1.21 m in diameter, spaced apart by approx- imately 17% of the rotor radius. The rotors are designed to operate at speeds up to 2800 RPM. The speed is fixed and a cell-phone class processor hosting the vision-based for the duration of flight, depending primarily on the current navigation functions. density, which is in the range 0.014–0.02 kg/m3. Control is via upper and lower swashplates, providing collective con- trol with a total range of 22◦, and cyclic control with a range 2.1 Demonstration Vehicle of 10◦, for each rotor. The vehicle used to demonstrate controlled flight in JPL’s Within the allowable density range, the vehicle is de- 25-ft Space Simulator is shown in Figure 4. It featured a signed to have the power and aerodynamic capability to full-scale rotor similar to the intended Mars vehicle. How- achieve thrust levels of at least 40% above hover thrust be- ever, since this vehicle was required to lift its own weight in fore the onset of stall. The helicopter would only fly in fa- Earth gravity, anything non-essential to the demonstration vorable weather, with wind velocities limited to 9 m/s hori- of controlled flight was left off the vehicle to reduce weight. zontally and 2 m/s vertically, with a maximum gust compo- Instead of onboard batteries, power was provided through nent of 3.5 m/s. Based on the forecasted weather, ground an electrical tether hanging below the vehicle. speed and climb/descent speeds would be limited such that maximum airspeed does not exceed 10 m/s horizontally and Unlike the proposed Mars vehicle, the flight demonstra- 3 m/s vertically. tion vehicle was only equipped with cyclic control on the lower rotor. This provides sufficient degrees of freedom for The gross vehicle weight is less than 1.8 kg, a substantial control, but results in reduced control authority and greater portion of which is taken up by the batteries. The batteries cross-axis coupling due to asymmetry between the rotors. provide energy for lasting up to 90 s, while also provid- For the demonstration vehicle, navigation was performed ing sufficient energy for non-flight operation of the onboard using a Vicon motion capture system, which determines electronics and night-time survival heating. The batteries both the position and attitude of the vehicle by tracking IR are rechargeable via a solar panel mounted on the non- retroreflective targets on the vehicle using external cam- rotating mast above the upper rotor. The batteries and other eras; this information was combined with an onboard IMU electronics are housed in a cube-like fuselage attached to in a filter. The navigation algorithm and all other flight con- the central mast, inside of which is a warm electronics box trol functionality was hosted on a ground computer. that is properly insulated and heated to protect against low night-time temperatures. Flights would be conducted based on a flight plan up- loaded from the ground, consisting of a series of waypoints. 3 Modeling Due to the many minutes of communication delay between Earth and Mars, each flight must be conducted with full The primary tool for studying flight dynamics for the pro- autonomy. For this purpose, onboard navigation is per- posed Mars Helicopter is the Helicopter Control Analysis formed using a combination of a small, MEMS-based in- Tool (HeliCAT), which was developed specifically for this ertial measurement unit (IMU), a low-resolution downward- purpose using the Darts/Dshell multibody simulation frame- looking camera, and a laser rangefinder. The IMU, which work developed at JPL [see, e.g., 8]. The strength of this measures accelerations and angular rates, is used for prop- tool is that it is tailored toward needs in the guidance, navi- agation of the vehicle state from one time step to the next. gation, and control area, with features including: The camera is used together with the laser rangefinder to determine the height above ground and the translational • detailed modeling of actuators and sensors, including velocity; this information is fused with the IMU solution. camera imaging The onboard computation platform consists of a radiation- tolerant FPGA; a dual-redundant automotive-class micro- • modeling of ground contact dynamics, including varied controller hosting the most critical flight control functions; terrain and surface properties

© 2017. All rights reserved. • modeling of ground support equipment, such as gim- bals and force-torque sensors, for validation of the sys- 1.2

tem identification approach 1

• flight software integration 0.8

• 3D visualization 0.6

In addition to generating models for control analysis, Heli- 0.4 CAT can be used to execute simulations of end-to-end mis- sions with flight software in the loop. Lift coefficient 0.2

The aircraft components are modeled in HeliCAT using 0 rigid bodies connected by hinges. Other rotorcraft modeling tools offer higher-fidelity modeling of flexible bodies; these -0.2 include CAMRAD II, a comprehensive analysis tool for ro- -0.4 torcraft that is widely used across NASA and in industry. -5 0 5 10 15 CAMRAD II is used in the Mars Helicopter project to validate Angle of attack (deg) certain properties of HeliCAT and to perform higher-fidelity analysis related to thrust and power performance. Figure 5: Typical plot of the lift coefficient as a function of angle-of- attack. 3.1 HeliCAT Modeling

The main bodies in the HeliCAT model are the fuselage, the the rotor disk. The magnitudes depend on the generated mast, the blades, and the legs. The connection thrust and aerodynamic moments, as well as the vertical of the blades to the mast is configurable as a sequence of and horizontal advance ratios. We use one model per ro- hinges with different properties, to represent different rotor tor, but simulate interference by increasing the vertical ad- designs. vance ratio used in each model by a fraction of the uniform inflow component of the other rotor. Mars Helicopter experi- 3.1.1 Aerodynamic Forces mental results to date indicate that a one-way coupling, with the lower rotor seeing the entire uniform component of the The aerodynamic forces on the blade are modeled by divid- upper rotor, is a reasonable approximation. To account for ing the blade into a large number of slices with individual ground effect, the uniform inflow component is restricted as span, chord, twist, and sweep. Associated with each el- a function of the distance above the terrain [see 6, Ch. 4.8]. ement is an airfoil table that tabulates the nondimensional Inflow modeling is known to be challenging; models of this coefficients of lift, drag, and pitch moment as a function of type are semi-empirical and have been developed for large- angle-of-attack and Mach number. The airfoil tables were scale, single-rotor helicopters on Earth. Thus, significant generated by AeroVironment from two-dimensional CFD inaccuracies are expected when applying the same models analysis of the specific blade section, using ANSYS Fluent. to small-scale co-axial helicopters in Mars atmosphere. As At each time step, forces and moments for each element are described in Section 8, modifications must be implemented computed based on interpolating the airfoil data. Despite to account for effects observed in experiments. the low Reynolds number regime, the lift, drag, and mo- ment characteristics are not dramatically different from typ- ical characteristics at Earth densities; crucially for the flight dynamics, the lift curves exhibit a linear region as seen in 3.2 Trim and Linearization Figure 5. The two-dimensional airfoil tables only account for flow For control analysis and design, it is desirable to obtain lin- perpendicular to the quarter-chord of the blade. An estimate ear time-invariant (LTI) models of the flight dynamics, as of radial drag effects due to yawed flow is included, based this enables the use of a wide variety of frequency-domain on the same airfoil tables [see 6, Ch. 6.22]. Tip losses are design and analysis tools. For helicopters, this is compli- modeled by eliminating the lift and moment contribution (but cated by the periodic nature of the dynamics. Equilibrium not the drag) of the outermost portion of the blade. Unsteady in steady-state flight is characterized by a periodic trajec- aerodynamic effects are also included, based on the angular tory, and linearizations along the equilibrium trajectory can velocity and acceleration of the blade [see 6, Ch. 10.6]. be highly time-varying. Fuselage drag is modeled by a simple blunt-body drag Similar to CAMRAD II, HeliCAT can be used to compute model with no angle-of-attack dependency; due to the low the periodic equilibrium trajectory for a given trim condition, density, this has little impact on the flight dynamics. and to output linearized models at a finite set of uniformly spaced points along the trajectory. A simple way to ob- tain an LTI model is to average the resulting periodic model; 3.1.2 Inflow Modeling however, this method yields an inaccurate response for the Inflow is modeled based on the dynamic inflow model of Pe- Mars Helicopter model, due to the time-varying coupling be- ters and HaQuang [11]. This is a nonlinear model, intended tween the rotor and fuselage with two-bladed rotors. to be valid over a range of flight regimes, and therefore suit- We instead post-process periodic linear time-variant mod- able for simulations of end-to-end missions. Inflow is repre- els generated by HeliCAT using a method similar to Lopez sented as a uniform component plus a linear gradient over and Prasad [9]. The method extends the vehicle state by in-

© 2017. All rights reserved. cluding the coefficients of a truncated harmonic expansion: ∑n

(1) χ = χ0 + (χic cos(iΩt) + χis sin(iΩt)), i=1 where Ω is the rotor angular rate. The periodic system Figure 6: Illustration of blade flapping modeled by a central hinge with a spring. dynamics obtained from the linearized models can be de- scribed by truncated harmonic expansions of the system matrices: When cyclic pitch is applied to a helicopter blade, a pe- ( ) ∑n riodic change in lift is produced at the rotor frequency, with χ˙ = A0 + (Aic cos(iΩt) + Ais sin(iΩt)) χ maximum lift on one side of the rotor disk, and minimum lift on the opposite side. A mass-spring-damper subjected to a (2) ( i=1 ) ∑n periodic force will respond by settling into a periodic motion + B0 + (Bic cos(iΩt) + Bis sin(iΩt)) u. at the same frequency, but with a different phase than the in- i=1 put. Therefore, when cyclic control is applied to a helicopter We insert (1) into (2) and reduce products of frequencies to rotor, it will settle into a periodic flapping motion that reaches single frequencies through standard trigonometric identities, its maximum at some point later than the maximum cyclic discarding frequencies above nΩ. We extract into separate pitch. The flapping of the blade results in roll and pitch mo- equations those terms that depend on each linearly inde- ments that are in general a combination of two effects: a tilt pendent harmonic, resulting in equations on the form of the thrust vector, which results in moments proportional to the height of the rotor plane with respect to the center of ∑n gravity; and direct hub moments, due to resistance against ˙ − χ0 A00 χ0 + (A0kc χkc + A0ks χks) + B0u = 0, flapping at the hub. For a more in-depth discussion, see k=1 Padfield [10]. ( n ) ∑ The blue line in Figure 7 shows the magnitude and phase χ˙ − A χ + (A χ + A χ ) + B u c(iΩt) = 0, ic ic0 0 ickc kc icks ks ic response of the blade flap angle in response to cyclic input k=1 for a centrally hinged rotor with no hinge stiffness. For this ( ∑n ) type of rotor, the restoring forces are all due to centrifugal χ˙ − Ais0 χ + (Aiskc χ + Aisks χ ) + Bisu s(iΩt) = 0. is 0 kc ks stiffening, which places the natural frequency of the mass- k=1 spring-damper system exactly at the rotor frequency. As a Discarding the harmonic multiplier and combining yields a result, the peak of the blade flapping, and the moments due high-order LTI system. to blade flapping, occur 90◦ after the peak cyclic pitch input. The output from the system is treated similarly, by decom- If the hinge is offset from the center, or it has additional hinge posing the periodic system output as stiffness, the natural frequency will increase, and the phase ( ) ∑n lag between cyclic pitch input and blade flapping output will ◦ (3) y = C0 + (Cic cos(iΩt) + Cis sin(iΩt)) χ, be reduced, in some cases to as little as 30–40 [see 5]. i=1 inserting (1), and extracting the resulting constant term from 4.2 The Role of Aerodynamic Damping the right-hand side. The predominant source of damping of the flap motion is Empirically, an expansion up to n = 4 is necessary and typically aerodynamic: as the blade flaps up or down, the sufficient for representing the Mars Helicopter frequency re- angle-of-attack changes as a function of the flap rate; this sponse with good accuracy. produces a change in lift opposing the flap rate. For the simplified blade model with a central hinge, if the blade is 4 Flap Dynamics additionally assumed to have a straight chord and a stan- dard linear lift model is used, the nondimensional damping The flight dynamics of a helicopter are highly influenced by is given by the dynamics of blade flapping. For a helicopter on Mars, γ the blade flapping dynamics diverges in important ways from (4) ζ = , 16ω that of Earth helicopters, due to to the low density of the 0 atmosphere. where ω0 is the flap frequency nondimensionalized by the rotor frequency; and γ is the blade Lock number, a measure of the ratio of aerodynamic to inertial forces, given by 4.1 Fundamentals of Blade Flapping ρcC R4 When studying how blade flapping is affected by the low (5) γ = lα . density of the Martian atmosphere, it is useful to start with I the simplest-possible model of the blade, as a rod rotat- Here, ρ is the air density, c is the chord length, Clα is the lift ing about a flap hinge with a spring, as shown in Figure curve slope of the blade airfoil, R is the rotor radius, and I is 6. Restoring moments about the flap hinge are present the blade inertia about the flap hinge. Blade Lock numbers due to a combination of centrifugal stiffening and moments for Earth helicopters are typically somewhere between 4 and from the spring, and damping is present due to aerody- 16. The response illustrated by the blue line in Figure 7 namic forces. The blade therefore acts like a classical mass- has a damping ratio of 25%, corresponding to a blade Lock spring-damper system. number of 4 and ω0 = 1.

© 2017. All rights reserved. o ld ign tanniesoa frequency nondimensional a Thus, at blade. ringing the blade of azimuth a current for the and angle flap the of fuselage. helicopter the to rotor attached observer the an in by is the interest in the the dynamics helicopter, from the studied controlling when When blade the represents the section of last the motion in studied dynamics Dy- flap The Helicopter the on Modes Flap of Effect 4.3 whole. a as helicopter the the in of modes response damped attitude poorly as itself ul- manifests steady-state; this to timately converge to longer takes con- spring-damper of point-of-view the the is from pitch trol and concerning roll More the produce to adjusting order moments. by in inputs for the cyclic the compensated of of easily clocking properties is it controllability as the helicopter, for important mentally frequency. natural the increase near- to to stiffened is drop rotor quickly the if will zero and frequency, natural the around o0 damping to the with 2 response to the The shows reduced dynamics. 7 flap Figure the in of line damping green the on effect in dramatic roughly a for has falls Indeed, it 0 design Mars. between range baseline on Helicopter reduced Mars significantly the be to it expect frequency. rotor the marks number line Lock red blade with stiffness. response number additional Lock Mars blade hypothetical no with with response blade Earth typical hinged Blue: centrally a for input, pitch 7: Figure h oet ntennoaigfaeaeafnto both function a are frame nonrotating the in moments The funda- not is steady-state at angle phase azimuthal The ic h ld oknme clswt est,w can we density, with scales number Lock blade the Since . 3 ti la httepaeagei xrml sensitive extremely is angle phase the that clear is It 33. Phase (deg) namics -180 -135 Magnitude (dB) -90 -45 -50 -40 -30 -20 -10 10 0 0 10 antd n hs fbaefa nl epneto response angle flap blade of phase and Magnitude transient 1 % orsodn oaLc ubrreduction number Lock a to corresponding , . non-rotating n 0 and 3 lprsos.Apol apdmass- damped poorly A response. flap Blade flapresponsetopitchinput . ,dpnigo h est,which density, the on depending 6, Frequency (Hz) rm—hti,wa sseen is what is, frame—that rotating γ γ = = 07 l ihsreserved. rights All 2017. © 0 .Green: 4. . ω 10 3 The 33. frame. 0 2 the , oal o uhaslto nteMr eiotr This Helicopter. Mars the on solution fa- a not such are circumstances for the vorable fil- stabilization, notch phase of or techniques tering using resonances, around” “work to to leading resonance, the of amplify response instability. may the system phase, frequency. control and same the magnitude the at the input on control actuator Depending feedback an the apply frequency, will vehi- resonant to system the response at In destructive motion itself. for cle system potential control are the the 8 with to Figure coupling due in control, shown for ones the problematic like modes damped Poorly Con- for Dynamics Flap of Consequences 5 flap advancing and regressing The visible. clearly point. are stud- modes early was an that partic- dynamics at a helicopter for ied for Mars angle, the pitch of to disk) version rotor ular the (peak- of cyclic rear the cosine at the from ing shows function 8 transfer Figure the up of fuselage. freeing magnitude the the to of of due condition dynamics frequency boundary flight shifted the a the These at in vehicle, modes of mode. free-flying as level flap up low non-rotating show the the ultimately inherit by they exhibited that damping show coordinates, plane the approximately damped at frame, poorly rotating single, frequency a at that mode phenomenon the illustrates where (6) form the of be will axis fixed a about moment are modes flap dynam- advancing Helicopter and visible. Mars regressing clearly the damped of poorly version The particular ics. a for input, cyclic 8: Figure lhuhi oecssi spsil o oto system control a for possible is it cases some in Although tip-path- and coning using modes, these of study formal A

-80 -60 -40 Magnitude-20 (dB) 20 40 regressing 0 rladVhceDesign Vehicle and trol A ψ sin stebaeaiuhand azimuth blade the is antd fpthagefeunyrsos ocosine to response frequency angle pitch of Magnitude ( ω 0 ψ 10 and ) -1 sin advancing ( ω ψ 0 + ie iet w oe ntenon- the in modes two to rise gives φ = ) Frequency (Hz) Bode Diagram 10 0 − modes. 2 1 2 1 A ω A cos φ 0 cos − sapaeage This angle. phase a is (( and 1 (( ω 10 ω 0 1 0 − + ω 1 ) 0 1 ψ ) + ψ − ,called 1, + φ φ ) 10 ) , 2 is in part due to the difficulty of accurately identifying the gust sensitivity and because it negatively affects the stability high-frequency dynamics; and in part because the high- properties of the vehicle. frequency dynamics, once other sources of flexibility are in- cluded, become far more complicated than suggested by Figure 8. 7 Low-Frequency Dynamics A robust solution is is to ensure that the resonant frequen- cies are all well above the bandwidth of the control system, A benefit of the stiff rotor design is that the linearized flight thereby preventing destructive control system interactions. dynamics can be accurately described by a low-order model This is achieved through the mechanical design of the he- with only vehicle velocities, attitude, and angular rates as licopter by making the blades and hub for the Mars Heli- states, in the frequency range relevant for control. This copter unusually stiff, with a rotating flap frequency of ap- model can be written as proximately 80–90 Hz. Together with other stiffness require- (7) x˙ = Ax + Bu, ments for the mast and landing gear, all resonant modes in the flight dynamics are then moved to high frequencies. where x = [u; v; w; φ; θ; ψ; p; q; r] contains the body-frame velocities relative to an inertial wind frame, the attitude Euler angles, and the body-frame angular rates; and 6 Sensitivity to Edgewise Flow u = [θl0; θlc; θls; θu0; θuc; θus] contains the lower collec- tive, cosine cyclic, and sine cyclic, followed by the upper Helicopters typically react to edgewise flow by producing a collective, cosine cyclic, and sine cyclic. We use the pitch moment away from the wind. In a traditional helicopter, convention that the x axis points forward on the vehi- this is primarily due to the differences in lift on the advancing cle; the y axis points to the right; and the z axis points and retreating side of the rotor disk, which results in an ef- down. The A and B matrices are written in standard form as fect similar to a cyclic control input peaking on the advancing   − ¯ − side. When the moments generated due to periodic inputs Xu Xv Xw 0 g cθ 0 Xp Xq w¯ Xr + ¯v  ¯ ¯  Yu Yv Yw g cφ¯ cθ −g sφ¯ sθ 0 Yp + w¯ Yq Yr − u¯ have a significant phase lag, as is the case for Earth heli-  ¯ ¯  Zu Zv Zw −g sφ¯ cθ −g cφ¯ sθ 0 Zp − ¯v Zq + u¯ Zr    copters, the result is a pitch moment away from the wind.  0 0 0 0 0 0 1 sφ¯ tθ¯ cφ¯ tθ¯    As explained in Section 4.2, the azimuthal phase lag in re- A =  0 0 0 0 0 0 0 cφ¯ −sφ¯     0 0 0 0 0 0 0 sφ¯/cθ¯ cφ¯/cθ¯ sponse to periodic inputs is very small for the Mars Heli-    Lu Lv Lw 0 0 0 Lp Lq Lr  copter; thus, lift differences on the advancing and retreating   Mu Mv Mw 0 0 0 Mp Mq Mr side will mostly result in a roll moments that are canceled Nu Nv Nw 0 0 0 Np Nq Nr out by an opposing moments from the other rotor. and However, another effect, due to inflow gradients across   the rotor disk, comes into play, with much the same result. XL0 XLC XLS XU0 XUC XUS   When the wake angle changes from the vertical due to the YL0 YLC YLS YU0 YUC YUS    vehicle moving relative to the surrounding air, a longitudinal ZL0 ZLC ZLS ZU0 ZUC ZUS    inflow gradient develops, resulting in more inflow at the rear  0 0 0 0 0 0    of the disk and less at the front. This again results in a peri- B =  0 0 0 0 0 0  .   odic change in lift, peaking at the front of the rotor disk. Due  0 0 0 0 0 0    to the small azimuthal phase lag of the moments generated  LL0 LLC LLS LU0 LUC LUS    due to periodic inputs, the result is a pitch moment away ML0 MLC MLS MU0 MUC MUS from the wind. Rather than cancel each other, the moments NL0 NLC NLS NU0 NUC NUS from the two rotors act in concert. We have denoted by [u¯; ¯v; w¯] the body-frame trim velocities; by φ¯ and θ¯ the trim roll and pitch attitudes; and by g the ac- 6.1 Drawback of Stiff Rotor Design celeration of gravity. The remaining components of A and B are stability and control derivatives that capture sensitivities The stiff rotor design exacerbates the vehicle’s sensitivity to of the vehicle motion to the states and control inputs. edgewise flow, through several different mechanisms:

• by reducing the azimuthal phase angle of moments in 7.1 Model Reduction response to periodic inputs, the moments generated by the upper and lower rotor become more closely aligned, Via a custom optimization process, low-order models on the creating a larger additive moment form (7) can be generated to match the frequency response of a full-order model derived from the process outlined in • the stiff rotor produces larger hub moments in response Section 3.2. The process can be summarized as follows: to edgewise flow, resulting in a faster transient re- We set up the full-order linear system so that its inputs match sponse those of the reduced-order system, and its outputs match the states of the reduced-order system. The full-order sys- • the stiff rotor reduces the apparent inertia of the coaxial tem with these inputs and outputs can be written in transfer helicopter—a phenomenon further discussed in Sec- function form as tion 7.2—which also results in a faster transient re- sponse. (8) x(s) = H(s)u(s).

As will be discussed in Section 7.3, increased sensitivity We select a set of evaluation frequencies ωi, i = 1,..., n, to edgewise flow is undesirable both because of increased logarithmically spaced over the frequency range of interest.

© 2017. All rights reserved. For each frequency ωi, we compute the steady-state fre- Bode Diagram quency response Hi := H(jωi). For each input component k, and for each state com- Full order 20 Reduced order ponent ℓ, we know that the input uk = sin(ωit) yields the 0 steady-state response xℓ = |hiℓk|sin(ωit + ∠hiℓk), where hiℓk is element (ℓ, k) of Hi. We furthermore know that the steady- -20 ˙ | | state response of the time derivative is xℓ = ωi hiℓk cos(ωit+ -40 ∠hiℓk). We now want to find A and B such that the same is true for the reduced-order system (7). This would require -60 -80

that the following hold: Magnitude (dB)

  -100 |hi11| c(ωit + ∠hi11) · · · |hi16| c(ωit + ∠hi16)  . . .  -120 ωi  . .. .  = . . -140 |hi91| c(ωit + ∠hi91) · · · |hi96| c(ωit + ∠hi96) 10-2 100 102   Frequency (Hz) (9) |hi11| s(ωit + ∠hi11) · · · |hi16| s(ωit + ∠hi16)    . . .  A . .. . Figure 9: Comparison of the full-order and reduced-order linear |hi91| s(ωit + ∠hi91) · · · |hi96| s(ωit + ∠hi96) models, for an up-to-date model of the Mars Helicopter. The com- +B s(ωit). plicated dynamics at high frequency is a result of poorly damped rotor dynamics, combined with flexibility of the mast and the land- The above set of 9 · 6 = 54, equations, which is continuous ing gear. in t, can be reduced to two sets of equations independent of t by evaluating at t = 0 and t = π/(2ωi). This involves no 2 loss of information, since (9) can be written for any other t Iav = Ic + 2Ib = 0.024 kg·m . We apply a constant torque as a linear combination of those two sets of equations. to this body of magnitude τx = 0.1 Nm about the x axis, By combining the two sets of equations extracted for each as the only external force in the system. Figure 10 shows frequency ωi, we have obtained a total of 54 · 2 · n = 108n the simulated angular velocities of the central body in re- linear equations with real coefficients. The unknowns are sponse to this torque. When the high-frequency vibrations those elements of A and B that are not known a priori. We are averaged out, the system exhibits a linear increase in x now estimate the unknowns by performing a least-squares axis angular velocity at a rate of 4.2 rad/s2, which matches 2 fit, thus yielding the A and B matrices for the reduced-order the formula τx/Iav = 4.2 rad/s . Thus, we conclude that the model. low-frequency response is indeed as hypothesized in the Figure 9 shows the magnitude of the full-order and previous paragraph. reduced-order frequency response of the transfer function Next consider the same system, but with the rotors free from cosine cyclic input to the helicopter pitch angle, for an to flap about a central hinge. The hinge is stiffened with a up-to-date model of the Mars Helicopter in hover. It is no- spring with a spring rate of 500 Nm/rad, which yields a rotat- table that the low-order model matches the high-order model ing flap frequency of 91 Hz when centrifugal stiffening is ac- well up to frequencies of approximately 10 Hz. This is ad- counted for. Figure 11 shows the simulated response. The equate for relying on the low-order formulation for control angular acceleration is still about the x axis, but the mag- analysis and design. nitude is reduced to 3.5 rad/s2—that is, the system still re- sponds similar to a rigid body, but it responds as though a smaller torque of approximately 0.08 Nm is being applied. 7.2 Apparent Inertia To understand this phenomenon, we need to locate the A rigid body with no angular momentum, when subjected to “missing” 0.02 Nm torque that is being applied to the sys- a torque of magnitude τ about a principal axis, responds with tem, but which does not manifest itself in the gross vehi- an angular acceleration of magnitude τ/I about the same cle motion. To this end, consider Figure 12, where the tip- axis, where I is the moment of inertia about the relevant axis. path-plane of each rotor is plotted for the same simulation. For a coaxial helicopter with stiff, counter-rotating rotors, it The tip-path-planes exhibit a linearly growing tilt about the would seem reasonable to expect a low-frequency behav- y axis in response to the x axis input, in opposite direc- ior consistent with that of a rigid body, with an inertia cor- tions. The rate of change has a magnitude of approximately responding to the average inertia of the system helicopter. β˙ = 0.01 rad/s when high-frequency vibrations are aver- aged out. Using standard equations for gyroscope motion, To investigate this hypothesis, consider a simple exam- we find that the torque required to effect this linear growth ˙ ˙ ple system consisting of a rigid central body attached to two is βH = 2βΩIb ≈ 0.01 Nm for each rotor, where H denotes perfectly stiff, 2-bladed counter-rotating rotors. The cen- the magnitude of the angular momentum for each rotor. This tral body has its center of mass at the origin and has a accounts for the missing torque. 2 diagonal inertia matrix of magnitude Ic = 0.02 kg·m in The phenomenon observed in this example is due to gy- all three axes. The rotors are both centered at the origin roscopic effect on the two rotors, which causes the rotors and rotate about the z axis at a rate of Ω = 272 rad/s, to realign themselves with opposing tilts about the y axis in and each blade has the inertia of a simple rod—zero about response to a torque about the x axis. While the rigid rotor 2 the symmetry axis and Ib = 0.002 kg·m in the other two head prevents any reorientation from happening, the flexible axes. The average inertia over one rotation is therefore rotor head allows for a limited amount of reorientation—even

© 2017. All rights reserved. 30 0.05 x rate Rotor 1 y axis y rate 0.04 Rotor 1 x axis z rate Rotor 2 y axis 25 Rigid-body x rate Rotor 2 x axis 0.03

0.02 20 0.01

15 0

-0.01 10

Angular rates (deg/s) Tip path planes (deg) -0.02

-0.03 5 -0.04

0 -0.05 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 Time (s) Time (s)

Figure 10: Angular velocity response to constant torque, for the Figure 12: Tip-path-plane response to constant torque, for the sim- simplified example system with a perfectly rigid rotor. The dashed plified example system with a flapping rotor. black line shows the response expected from a perfect rigid body with an inertia corresponding to the average inertia of the rotating system. angular momentum is · (11) Htotal = Iactual ωx + H1 βy1 + H2 βy2, 30 x rate where H1 and H2 are the angular momentums of the gyro- y rate z rate scopes, and β and β are the gyroscope tilts about the 25 Rigid-body x rate y1 y2 central body’s y axis. For each gyroscope to maintain a steady-state rate about the x axis, it must be acted upon by 20 a y axis torque −H{1,2}ωx, which must necessarily be equal − to the spring torque kβy{1,2}, since this is the only external 15 torque on the gyroscope. Combining these expressions, we have 10

Angular rates (deg/s) H{ }ωx (12) β = 1,2 . y{1,2} k 5 Inserting into (11), we obtain

0 2H2ω 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 (13) H = I · ω + x . Time (s) total actual x k

If we now compute Iapparent = Htotal/ωx, we obtain (10). Figure 11: Angular velocity response to constant torque, for the The formula (10) can be used to approximate the effect simplified example system with a flapping rotor. The dashed black for a helicopter with discrete blades, by replacing Iactual with line shows the response expected from a perfect rigid body with an the average inertia, and k with the spring constant of each inertia corresponding to the average inertia of the rotating system. blade, multiplied by N/2, where N is the number of blades on each rotor. Applied to the example in the last section, we · 2 for very stiff rotors—thereby “absorbing” some of the torque obtain Iapparent = 0.029 kg m , which is consistent with the applied to the system and giving it a larger apparent inertia. observed response.

7.3 Open-Loop Stability 7.2.1 A Formula for Apparent Inertia Unlike most vehicles, helicopters are almost always unsta- By modeling the system as two counter-rotating gyroscopes ble in open loop, with poles in the open right-half complex on a central body, and calculating the DC gain from a torque plane. This is no different for helicopters on Mars, as indi- input to the angular accelerations, the following formula can cated in Figure 13, which shows the open-loop poles of the be derived for the apparent inertia: hover dynamics for an up-to-date model of the Mars Heli- 2H2 copter. The dynamics exhibit four stable subsidence modes (10) Iapparent = Iactual + , in the open left-half complex plane; one marginally stable k mode at the origin; and four unstable modes in the open where Iactual is the total x/y inertia of the system, and k is the right-half complex plane. The mode at the origin is due to torsional spring constant between each gyroscope and the the yaw angle being included in the state vector. The modes central body. in the right-half complex plane are longitudinal and lateral A very simple argument can be formulated as follows: for phugoid modes that result from a coupling between the at- a steady-state angular rate ωx about the x axis, the total titude and horizontal speed states.

© 2017. All rights reserved. 2 2

1.5 1.5

1 1

0.5 0.5

0 0 Imaginary Imaginary -0.5 -0.5

-1 -1

-1.5 -1.5

-2 -2 -2 -1.5 -1 -0.5 0 0.5 1 1.5 -4 -3 -2 -1 0 1 2 3 4 5 Real Real

Figure 13: Poles of the low-frequency dynamics of the demonstra- Figure 14: Root loci of the helicopter dynamics, transitioning from tion vehicle. hover to forward flight at 10 m/s airspeed. The blue and red ar- rows indicate the evolution of the longitudinal and lateral phugoids, respectively. The pole moving further out into the right-half com- To see what drives these modes, consider the longitudi- plex plane corresponds to the pitch-heave instability that develops nal dynamics only, restricted to the pitch angle, pitch rate, in forward flight. and longitudinal velocity. Assume furthermore that Xu, Xq, and Mq are zero, which is close to the truth for the Mars He- a positive trim longitudinal velocity u¯. In response to a posi- licopter. At hover with zero trim angles, the resulting system tive perturbation in w (i.e., a downward movement), the he- matrix is given by licopter pitches nose-up due to the positive M , resulting   w 0 −g 0 in a positive pitch rate. Due to the kinematic coupling be-   tween pitch rate and vertical velocity, captured by the ele- Al = 0 0 1 . ment Zq +u¯ in the A matrix, w increases further, thus making Mu 0 0 the effect self-amplifying. The stability derivative Mu represents the sensitivity of the For the Mars Helicopter flight dynamics model, the posi- helicopter’s pitch rate to longitudinal speed; or equivalently, tive Mw in forward flight is primarily driven by inflow effects. the sensitivity to a gust from the front. It is positive, because In trim forward flight, the wake angle is deflected from the increased longitudinal velocity results in a nose-up moment. vertical, resulting in an inflow gradient whose effect is coun- 3 The characteristic equation√ of Al is given by√λ +√Mug = 0, teracted by the cyclic trim settings. A positive perturbation − 3 1  3 with solutions λ1 = Mug, λ{2,3} = 2 (1 3j) Mug. It in w reduces the net inflow through the rotor, thus increasing is clear that the frequency of the unstable poles increases the wake angle. This results in an increased inflow gradient, with both Mu and g. We can therefore make the following which gives rise to a nose-up pitch moment. observations: • As discussed in Section 6, the high rotor stiffness re- 8 System Identification sults in increased sensitivity to edgewise flow, which manifests itself as an increase in the magnitude of Mu. The identification of the helicopter bare-airframe system This results in an increase in the frequency of the un- dynamics is crucial to the design of a stable flight control stable modes. system. Despite the extensive modeling and simulation • The unstable modes for a helicopter operating in Mars effort, testing is needed to fully understand the vehicle be- gravity will have reduced frequency compared to Earth. havior, especially given the lack of prior work on helicopter dynamics in Martian conditions. A system identification The unstable nature of the open-loop dynamics is highly sig- program has therefore been developed to identify the actual nificant in the context of control design, as it imposes funda- dynamics of the helicopter. mental limitations on the achievable stability margins, which become more severe as the frequency of the instability in- A unique challenge of flight dynamics testing of a Mars creases [see 14]. Helicopter is the difficulty of replicating the Martian condi- tions; in particular, low density conditions cannot be com- 7.4 Forward Flight bined with emulation of Martian gravity in a practical way. While helicopter mass can be offloaded to match the weight The dynamics of the helicopter changes as it moves from on Mars, this cannot be achieved without changing the cen- hover to forward flight. In particular, the helicopter exhibits ter of mass and inertia of the vehicle. Further limitations in- the classical trait of developing a pitch-heave instability, as clude the availability of a test facility with sufficient space for seen from the root-loci in Figure 14. end-to-end testing, and the complication of replicating the The pitch-heave instability is driven by the emergence of full range of environmental disturbances (wind and gusts). a positive Mw stability derivative. Consider forward flight at Given these limitations, a “piecewise” system identification

© 2017. All rights reserved. approach is adopted, in which individual parameters rele- In order to maintain stability on the gimbal, a low- vant to the bare-airframe dynamics (i.e., control derivatives, bandwidth controller was implemented to maintain attitude speed stability derivatives, attitude damping derivatives, in- while minimizing control input cross-coupling. A tether with ertias, etc.) are targeted via specific tests designed to be a pre-loaded spring was used to restrain the vehicle during performed in the confined space of a vacuum chamber. spin-up and spin-down, and to partially restrain the vehicle during some of the tests. This tether had the effect of in- troducing a restoring moment as a function of the attitude, 8.1 System Identification for the Demonstration which had to be accounted for in the identification process. Vehicle Different system identification methods, including the The system identification process for the demonstration ve- time-domain examination of the steady-state response to hicle focused on the hover condition only, using a process doublet inputs, and frequency-domain methods for the anal- based on three test configurations: ysis of the system response to sinusoidal frequency sweeps were employed. The latter frequency-response analysis 1. Locked-down configuration: the helicopter was made extensive use of the methods implemented in the mounted on a force-torque sensor. software package Comprehensive Identification from Fre- quency Responses (CIFER), developed at Ames Research 2. Swinging-arm configuration: the helicopter was Center [see 16]. mounted on a force-torque sensor at the end of a swinging arm. 8.1.1 System model 3. Gimbal configuration: the helicopter was mounted on a The model to be identified was defined by a slightly modified gimbal with roll and pitch degrees of freedom. form of (7) that is considered valid for the hover condition: In the locked-down and gimbal configurations, the helicopter controls were exercised with different inputs. On the swing- (14) Mx˙ = Fx + Gu + rG, ing arm, the arm motion was prescribed to emulate a de- where sired translational motion, while approximately fixing the   rotational motion (the swinging motion inherently imparts m 0 0 0 0 0 0 0 0   some angular motion).  0 m 0 0 0 0 0 0 0    The first tests were conducted with the helicopter locked  0 0 m 0 0 0 0 0 0    down. In this configuration, the rotational degrees of free-  0 0 0 1 0 0 0 0 0    dom of the vehicle were constrained, and the controls were (15) M =  0 0 0 0 1 0 0 0 0    subjected to:  0 0 0 0 0 1 0 0 0     0 0 0 0 0 0 Ixx Ixy 0  • high-frequency sweeps: designed to verify available   0 0 0 0 0 0 Iyx Iyy 0 actuator control bandwidth and that undesirable high- 0 0 0 0 0 0 0 0 Izz frequency rotor modes are not within the frequency range desired for control; and is the “mass matrix” of the helicopter. The total vehicle mass m and the yaw inertia Izz are known parameters and can be • moderate-duration doublets: designed to identify treated as such in the system identification problem. How- the quasi-steady vehicle control (force and moment) ever, because of the apparent inertia effect, the roll and pitch derivatives. inertias were treated as unknown parameters to be identi- Force and torque reactions on the force-torque sensor were fied. The remaining matrices in (14) are measured and used for estimation of the derivatives with   Xu Xv 0 0 −mg 0 Xp Xq 0 respect to the control inputs. The apparent roll and pitch   Yu Yv 0 mg 0 0 Yp Yq 0  inertia of the vehicle was unknown at this stage; thus the    0 0 Zw 0 0 0 0 0 Zr  absolute magnitude of the control derivatives in roll and pitch    0 0 0 0 0 0 1 0 0  could not yet be determined.   F =  0 0 0 0 0 0 0 1 0  On the swinging arm, the vehicle was swung back and    0 0 0 0 0 0 0 0 1  forth to generate edgewise flow over the rotor; and to gen-    Lu Lv 0 0 0 0 Lp Lq 0  erate flow perturbations through the rotor by changing the M M 0 0 0 0 M M 0  mounting orientation on the arm. The resulting forces and u v p q 0 0 N 0 0 0 0 0 N moments were measured to determine the speed stability w r derivatives. As with the control derivatives, roll and pitch and moment speed derivatives could not yet be positively iden-   0 XLC XLS 0 tified because of the unknown apparent inertia.    0 YLC YLS 0  Finally, control input frequency sweeps were applied to   ZS0 0 0 ZA0  the helicopter on the gimbal, which allowed the excitation    0 0 0 0  of the roll and pitch rates required for the system identifi-   G =  0 0 0 0  . cation of the bare-airframe roll and pitch dynamics. Among    0 0 0 0  the system parameters identified was the apparent inertia in    0 LLC LLS 0  roll and pitch, which provided the missing piece of informa-   tion necessary to determine the control and speed stability 0 MLC MLS 0 derivatives in roll and pitch. NS0 0 0 NA0

© 2017. All rights reserved. The corresponding state vector is arranged as x = 1 [u; v; w; φ; θ; ψ; p; q; r]. The input vector is defined as u = x measurement 0.8 y measurement [θs0; θlc; θls; θa0], where θs0 and θa0 are the symmetric and x model y model anti-symmetric collective components: 0.6

1 1 0.4 (16) θs0 = (θl0 + θu0) , θa0 = (θl0 − θu0) . 2 2 0.2

Swashplate servo-actuator dynamics were assumed to 0 be of a second-order nature, with a time delay: -0.2 Torques (Nm) − θ (s) ω2e τks -0.4 k = k . 2 2 -0.6 θkcmd (s) s + 2ζkωks + ωk It was ascertained early on that it would not be possible -0.8 -1 to experimentally measure the actual value of the bare- 335 340 345 350 355 360 Time (s) airframe control inputs θl0, θlc, θls, and θa0. Therefore, the inputs of the parametric model structure were actually de- fined in terms of the command input signals θs0cmd , θlccmd , θ , and θ . In doing so, the servo-actuator dynamic Figure 15: Comparison of measured and a priori modeled roll and lscmd a0cmd pitch torques in response to steps in the cyclic channels. Three- ◦ properties (ωk, ζk and τk) also become parameters of the second steps of 5 in cosine cyclic (positive, then negative) are system that need to be identified. Apart from the time de- followed by similar steps in sine cyclic. Note that cosine cyclic peaks lays, actuators were assumed to be identical, capturing the at the rear of the rotor disk (along negative x), and sine cyclic peaks global combined effect of the individual actuators. Specify- on the right-hand-side of the rotor disk (along positive y). The data ing different time delays allows for the problem to account has been filtered with a 2-Hz lowpass filter and a constant offset has been removed for better comparison. The less-clean data in the for unmodeled high-frequency dynamics in each axis which second-half of the plot are due to air recirculation in the chamber, may appear as an effective additional delay. which tended to manifest itself after 10–15 s in each test. The term rG on the right-hand side of (14) generically rep- resents the effect of the external forces and moments acting on the model, such as the gimbal or force-torque sensor re- roll and pitch damping. A value of KR = 2.9 was needed to actions. replicate the behavior in simulation models.

8.1.2 System Identification Findings Sensitivity to edgewise flow Data from the swinging-arm The system identification approach was successful in de- test, although noisy, indicates that the on-axis moments termining the key parameters of the system. Many of the generated due to edgewise flow, as captured by the stabil- smaller parameters could not be reliably identified due to ity derivatives Mu and Lv, were significantly larger than ex- noise and aerodynamic disturbances; however, this was ex- pected. This is most likely due to the development of a larger pected and is of little consequence because of their limited inflow gradient than predicted by the inflow model of Peters impact on the dynamics. and HaQuang [11]. By adding a tuning parameter to amplify The identified parameters matched analytical model pre- the sensitivity of the inflow gradient to edgewise flow by a dictions with reasonable accuracy, but notable differences factor of approximately 2.5, the results can be reproduced were found in certain areas, as discussed in the following in simulation. The increased sensitivity has a negative ef- paragraphs. Among these, only the first two have a signif- fect on the open-loop stability properties of the system (see icant impact on the stability and controllability properties of Section 7.3). the system. Unmodeled dynamics in thrust response Slow (0.4– Rotor damping Contrary to typical helicopter behavior, 0.5 Hz) unmodeled dynamics, resembling a highly damped rotor aerodynamic damping of the roll and pitch rates was pole-zero pair, were uncovered in the thrust response to found to be negative, meaning that roll and pitch rates are symmetric collective. The mode was sufficiently below the self-amplifying rather than decaying. This was reflected in control crossover frequencies and did not exhibit any per- the positive values of the identified on-axis roll and pitch sistent drop-off in phase, and as such was not considered damping derivatives Mq and Lp (values on the order of 0.4– problematic for control design. The source of this unmod- 0.45 rad/s). The likely cause is a dependence of the inflow eled dynamics is likely a coupling with the rotor RPM con- gradients on the roll and pitch rates of the rotor disk, which troller, rather than an aerodynamic effect. is not included in the inflow model of Peters and HaQuang [11] but can be found elsewhere [e.g., 6, Ch. 11.3] and is typically captured by a parameter denoted by KR. The inclu- Phasing of roll/pitch moments A priori modeling indi- sion of this effect increases the inflow over the rear side of cated that moments in response to cyclic would lead the the rotor disk for a nose-up pitch rate, which in turn creates applied cyclic by approximately 8◦, due to unsteady aerody- a larger positive moment. This effect on the inflow gradient namic effects from the angular velocity of the blade. Test typically changes the off-axis response to control inputs, but results instead indicated a lag of approximately 11◦ (see due to the peculiarities of the Mars Helicopter flap dynam- Figure 15). The lag cannot be explained by any reasonable ics (see Section 4), it instead has the effect of removing the change in the stiffness of the rotor; it is instead hypothesized

© 2017. All rights reserved. that it may be due to a small lag in the buildup of aerody- namic lift forces in response rapid angle-of-attack changes. The effect can be replicated in simulation models by the ad- dition of a first-order aerodynamic lag with a time constant of approximately 1.7 ms.

Thrust and yaw torque response to collective Moder- ate discrepancies in the steady-state thrust and torque re- sponses to symmetric and anti-symmetric collective step in- puts were found to be consistent with a momentum theory under-prediction of the uniform inflow component by approx- imately 20%.

Figure 16: The Mars Helicopter demonstration vehicle in hover, Apparent inertia The apparent inertia, as measured by inside JPL’s 25-ft Space Simulator. the relationship between the roll and pitch control deriva- tives and the moments measured on the force-torque sen- sor, was found to be larger than predicted based on simu- 0.5 lation, adding approximately 60% to the rigid-rotor case as % opposed to a predicted increase of approximately 50 . Al- 0 though the effect can be reproduced by lowering the rotor stiffness, measurements of the vehicle stiffness are not con- -0.5 sistent with this being the true underlying cause. We instead hypothesize that the increased apparent inertia is related to the distributed and high-order nature of the rotor flexing, -1

which is not captured by the single-hinge model used to pre- Position (m) dict the effect a priori. -1.5

-2 9 Free-Flight Demonstration

-2.5 Despite certain challenges related to the flight dynamics 115 120 125 130 135 140 145 150 155 160 165 Time (s) of the proposed Mars Helicopter, it is possible to design controllers with good stability margins and adequate perfor- mance within the intended flight envelope, provided laten- Figure 17: Estimated vehicle position (center of mass relative cies are kept small and the actuators have sufficiently high to ground frame, represented in ground frame) during free-flight bandwidth. Such a design was carried out for the demon- demonstration. Recall that positive z axis is down. stration vehicle described in Section 2.1, and the vehicle was flown successfully in JPL’s 25-ft Space Simulator, in 3 environmental testing (radiation, thermal, vibration, etc.). A CO2 at a density of 0.0175 kg/m . The flight was fully au- tonomous, consisting of takeoff, climb to an altitude of 2 m main point of interest in upcoming tests will be the identified at a rate of 1 m/s, hover for 30 s, descent at 0.5 m/s, and vehicle behavior in forward flight, which has been modeled landing. A picture of the flying vehicle can be seen in Fig- but not covered by actual tests to date. The helicopter dy- ure 16. The position of the vehicle during the demonstration namics in forward flight tends to exhibit faster instabilities flight is shown in Figure 17. During flight, the vehicle created and greater levels of cross-axis coupling. Analysis to date its own gusty weather within the chamber, which accounts nevertheless shows that adequate stability margins and per- for the variation in position during the hover portion. The formance can be achieved within the desired flight envelope. control design will be described in an upcoming paper [4]. References 10 Concluding Remarks [1] E. W. Aiken, R. A. Ormiston, and L. A. Young. Future di- As discussed in this paper, analysis of helicopter flight dy- rections in rotorcraft technology at Ames research cen- namics for Mars has revealed a number of significant dif- ter. In Proc. American Helicopter Society Annual Fo- ferences with typical Earth helicopter dynamics. Some of rum, Virginia Beach, VA, 2000. these pose challenges for closed-loop control. Nonethe- less, with the successful completion of a system identifica- [2] J. Balaram and P. T. Tokumaru. Rotorcrafts for Mars tion and free-flight demonstration campaign, the feasibility of exploration. In Proc. International Planetary Probe controlled helicopter flight in Martian atmosphere has been Workshop, Pasadena, CA, 2014. LPI Contribution No. established. This marks a significant step toward the ulti- 1795. mate goal of flying on Mars. At the current point in time, development is progressing [3] A. Datta, B. Roget, D. Griffiths, G. Pugliese, J. Sitara- on a flight-ready Mars vehicle, and tests are being prepared man, J. Bao, L. Liu, and O. Gamard. Design of the for further system identification and flight testing, as well as Martian autonomous rotary-wing vehicle. In Proc. AHS

© 2017. All rights reserved. Specialist Meeting on Aerodynamics, Acoustics, and [17] N. Tsuzuki, S. Sato, and T. Abe. Conceptual design Test and Evaluation, San Francisco, CA, 2002. and feasibility for a miniature Mars exploration rotor- craft. In Proc. International Congress of the Aeronauti- [4] H. F. Grip, D. P. Scharf, C. Malpica, W. Johnson, cal Sciences, 2004. M. Mandić, G. Singh, and L. Young. Guidance and con- trol for a Mars helicopter. Submitted to AIAA Guidance, Navigation, and Control Conference, 2018. [18] L. A. Young. Vertical lift – not just for terrestrial flight. In Proc. AHS/AIAA/RaeS/SAE International [5] D. H. Halley. ABC helicopter stability, control, and Conference, Arlington, VA, 2000. vibration evaluation on the Princeton Dynamic Model [19] L. A. Young and E. W. Aiken. Vertical lift planetary Track. In Proc. American Helicopter Society 29th An- aerial vehicles: Three planetary bodies and four con- nual Forum, Washington, DC, 1973. ceptual design cases. In Proc. European Rotorcraft Fo- [6] W. Johnson. Rotorcraft Aeromechanics. Cambridge rum, Moscow, Russia, 2001. University Press, 2013. [20] L. A. Young, R. T. N. Chen, E. W. Aiken, and G. A. [7] I. Kroo and P. Kunz. Development of the Mesicopter: Briggs. Design opportunities and challenges in the de- A miniature autonomous rotorcraft. In Proc. American velopment of vertical lift planetary aerial vehicles. In Helicopter Society Vertical Lift Aircraft Design Conf., Proc. American Helicopter Society International Verti- San Francisco, CA, 2000. cal Lift Aircraft Design Conference, San Francisco, CA, [8] C. Lim and A. Jain. Dshell++: A component based, 2000. reusable space system simulation framework. In Proc. Third IEEE International Conference on Space Mission [21] L. A. Young, E. W. Aiken, V. Gulick, R. Mancinelli, and Challenges for Information Technology, Pasadena, G. A. Briggs. Rotorcraft as Mars scouts. In Proc. IEEE CA, 2009. Aerospace Conf., Big Sky, MT, 2002. [9] M. Lopez and J. V. R. Prasad. Linear time invari- ant approximations of linear time periodic systems. In [22] L. A. Young, E. W. Aiken, and G. A. Briggs. Smart ro- Proc. European Rotorcraft Forum, Munich, Germany, torcraft field assistants for terrestrial and planetary sci- 2015. ence. In Proc. IEEE Aerospace Conference, Big Sky, [10] G. D. Padfield. Helicopter Flight Dynamics: The The- MT, 2004. ory and Application of Flying Qualitites and Simulation Modeling. AIAA, 1996. [23] L. A. Young, P. Lee, G. Briggs, and E. Aiken. Mars [11] D. A. Peters and N. HaQuang. Dynamic inflow for prac- rotorcraft: Possibilities, limitations, and implications for tical applications. J. American Helicopter Society, 33 human/robotic exploration. In Proc. IEEE Aerospace (4):64–68, 1988. Conference, Big Sky, MT, 2005.

[12] G. Savu and O. Trifu. Photovoltaic rotorcraft for Mars missions. In Proc. Joint Propulsion Conf. and Exhibit, [24] L. A. Young et al. Use of vertical lift planetary aerial San Diego, CA, 1995. vehicles for the exploration of Mars. In NASA Head- quarters and Lunar and Planetary Institute Workshop [13] H. Song and C. Underwood. A Mars VTOL aerobot on Mars Exploration Concepts, Houston, TX, 2000. – preliminary design, dynamics and control. In Proc. IEEE Aerospace Conference, 2007. [25] L. A. Young et al. Engineering studies into vertical lift [14] G. Stein. Respect the unstable. IEEE Control Systems planetary aerial vehicles. In Proc. AHS International Magazine, 23(4):12–25, 2003. Meeting on Advanced Rotorcraft Technology and Life Saving Activities, Utsunomiya, Tochigi, Japan, 2002. [15] B. Thompson. Full throttle to Mars. Rotor & Wing, Phillips Business Information, LLC, Potomac, MD, 2001. [26] L. A. Young et al. Experimental investigation and demonstration of rotary-wing technologies for flight in [16] M. B. Tischler and R. K. Remple. Aircraft and Rotorcraft the atmosphere of Mars. In Proc. American Heli- System Identification: Engineering Methods with Flight copter Society Annual Forum Proceedings, Montreal, Test Examples. AIAA, 2012. Canada, 2002.

© 2017. All rights reserved.