trajectory design with hybrid low-thrust propulsion systema).pdftrajectory design with hybrid...

8
Trajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali and Alessandro A. Quarta University of Pisa, 56122 Pisa, Italy DOI: 10.2514/1.22433 A novel mission concept based on a hybrid low-thrust propulsion system is proposed and discussed. A solar electric propulsion thruster is coupled with an auxiliary system providing an inverse square radial thrust. In this way the spacecraft is virtually subjected to a reduced gravitational solar force. The primary purpose of this paper is to quantify the impact of the reduced solar force on the propellant consumption for an interplanetary mission. To this end the steering law that minimizes the propellant consumption for a circle-to-circle rendezvous problem is found using an indirect approach. The hybrid system is compared with a conventional solar electric thruster in terms of payload mass fraction deliverable for a given mission. A tradeoff between payload size and trip time is established. Nomenclature A = sail area f m = auxiliary function [see Eq. (26)] H = Hamiltonian H 0 = Hamiltonian that explicitly depends on the control vector J = performance index m = mass r = sunspacecraft distance T = thrust t = time u = radial velocity u = control vector v = circumferential velocity v c = circular velocity v e = electric thruster exhaust velocity v esc = escape velocity x = ratio between the orbital radii = electric thruster direction angle = dimensionless sail loading ~ = weighted dimensionless sail loading V = total variation of velocity = ratio of the auxiliary system thrust force to the solar gravitational force = polar angle = adjoint variable = gravitational parameter = sail loading = electric thruster switching parameter a = solar sail switching parameter = sail optical efciency Subscripts a = auxiliary e = solar electric propulsion system f = nal max = maximum pay = payload prop = propellant sail = sail assembly 0 = initial * = critical = sun = Venus = Mars Superscripts = time derivative Introduction N EW mission concepts are increasingly considering the use of low-thrust propulsion for efcient navigation in deep space. The employment of low-thrust propulsion technologies is especially interesting for those missions requiring large changes in orbital energy [1,2]. The closely related problem of trajectory optimization is also highly attractive from a theoretical viewpoint due to the inherent difculties in nding optimal trajectories characterized by continuous thrust over long time periods. Several deep-space missions have been identied that can be performed using solar electric propulsion (SEP) so as to signicantly reduce the total mission costs. These missions include Venus surface sample return, Saturn ring observer, Titan explorer, Neptune orbiter, and various Mars sample return options [3]. Other low-thrust propulsion systems, such as solar sails and minimagnetospheric plasma propulsion, have been considered for primary propulsion systems in interplanetary missions, for example, see [4,5]. The choice among the different propulsion concepts is driven by various mission requirements and constraints, among which the total mission cost has a central role. The purpose of this paper is to investigate the potentialities introduced by integrating two different low-thrust propulsion concepts and to highlight the tradeoff between costs and ight time. The starting point of our analysis is that heliocentric transfers could benet from a reduced gravitational solar force acting on the spacecraft. In a recent paper McInnes [6] introduced the concept of generalized orbits, that is, orbits obtained through a modulated inverse square radial thrust. In essence, the idea consists of studying the motion of a spacecraft under the effect of a gravitational force whose magnitude is reduced with respect to the solar gravity. The size of this reduction, quantiable through a dimensionless parameter that will be referred to as , is, to some extent, variable and capable of being modulated as a function of time. With such a model different families of orbits can be generated, and the problem of calculating open escape orbits and transfers between circular coplanar orbits can be approached. The maximum orbit radius attainable using an inverse square radial thrust is r r 0 =1 max [6], where max is the maximum value of . There are at least two types of low-thrust propulsion, Received 13 January 2006; revision received 11 May 2006; accepted for publication 12 May 2006. Copyright © 2006 by Giovanni Mengali and Alessandro A. Quarta. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 0731-5090/07 $10.00 in correspondence with the CCC. Associate Professor, Department of Aerospace Engineering; g.mengali@ ing.unipi.it. Member AIAA. Research Assistant, Department of Aerospace Engineering; a.quarta@ ing.unipi.it. Member AIAA. JOURNAL OF GUIDANCE,CONTROL, AND DYNAMICS Vol. 30, No. 2, MarchApril 2007 419

Upload: others

Post on 22-Mar-2020

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

Trajectory Design with Hybrid Low-Thrust Propulsion System

Giovanni Mengali∗ and Alessandro A. Quarta†

University of Pisa, 56122 Pisa, Italy

DOI: 10.2514/1.22433

Anovelmission concept based on ahybrid low-thrust propulsion system is proposed anddiscussed.A solar electric

propulsion thruster is coupled with an auxiliary system providing an inverse square radial thrust. In this way the

spacecraft is virtually subjected to a reduced gravitational solar force. The primary purpose of this paper is to

quantify the impact of the reduced solar force on the propellant consumption for an interplanetary mission. To this

end the steering law that minimizes the propellant consumption for a circle-to-circle rendezvous problem is found

using an indirect approach. The hybrid system is compared with a conventional solar electric thruster in terms of

payload mass fraction deliverable for a given mission. A tradeoff between payload size and trip time is established.

Nomenclature

A = sail areafm = auxiliary function [see Eq. (26)]H = HamiltonianH0 = Hamiltonian that explicitly depends on the control

vectorJ = performance indexm = massr = sun–spacecraft distanceT = thrustt = timeu = radial velocityu = control vectorv = circumferential velocityvc = circular velocityve = electric thruster exhaust velocityvesc = escape velocityx = ratio between the orbital radii� = electric thruster direction angle�� = dimensionless sail loading~�� = weighted dimensionless sail loading�V = total variation of velocity� = ratio of the auxiliary system thrust force to the solar

gravitational force� = polar angle� = adjoint variable� = gravitational parameter� = sail loading� = electric thruster switching parameter�a = solar sail switching parameter� = sail optical efficiency

Subscripts

a = auxiliarye = solar electric propulsion systemf = finalmax = maximumpay = payloadprop = propellant

sail = sail assembly0 = initial* = critical� = sun♀ = Venus♂ = Mars

Superscripts

� = time derivative

Introduction

N EW mission concepts are increasingly considering the use oflow-thrust propulsion for efficient navigation in deep space.

The employment of low-thrust propulsion technologies is especiallyinteresting for those missions requiring large changes in orbitalenergy [1,2]. The closely related problem of trajectory optimizationis also highly attractive from a theoretical viewpoint due to theinherent difficulties in finding optimal trajectories characterized bycontinuous thrust over long time periods.

Several deep-space missions have been identified that can beperformed using solar electric propulsion (SEP) so as to significantlyreduce the total mission costs. These missions include Venus surfacesample return, Saturn ring observer, Titan explorer, Neptune orbiter,and various Mars sample return options [3]. Other low-thrustpropulsion systems, such as solar sails and minimagnetosphericplasma propulsion, have been considered for primary propulsionsystems in interplanetary missions, for example, see [4,5].

The choice among the different propulsion concepts is driven byvarious mission requirements and constraints, among which the totalmission cost has a central role. The purpose of this paper is toinvestigate the potentialities introduced by integrating two differentlow-thrust propulsion concepts and to highlight the tradeoff betweencosts and flight time.

The starting point of our analysis is that heliocentric transferscould benefit from a reduced gravitational solar force acting on thespacecraft. In a recent paper McInnes [6] introduced the concept ofgeneralized orbits, that is, orbits obtained through a modulatedinverse square radial thrust. In essence, the idea consists of studyingthe motion of a spacecraft under the effect of a gravitational forcewhose magnitude is reduced with respect to the solar gravity. Thesize of this reduction, quantifiable through a dimensionlessparameter that will be referred to as �, is, to some extent, variable andcapable of being modulated as a function of time. With such a modeldifferent families of orbits can be generated, and the problem ofcalculating open escape orbits and transfers between circularcoplanar orbits can be approached.

The maximum orbit radius attainable using an inverse squareradial thrust is r� r0=�1 � �max� [6], where �max is the maximumvalue of �. There are at least two types of low-thrust propulsion,

Received 13 January 2006; revision received 11 May 2006; accepted forpublication 12 May 2006. Copyright © 2006 by Giovanni Mengali andAlessandro A. Quarta. Published by the American Institute of AeronauticsandAstronautics, Inc., with permission. Copies of this papermay bemade forpersonal or internal use, on condition that the copier pay the $10.00 per-copyfee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers,MA 01923; include the code 0731-5090/07 $10.00 in correspondence withthe CCC.

∗Associate Professor, Department of Aerospace Engineering; [email protected]. Member AIAA.

†Research Assistant, Department of Aerospace Engineering; [email protected]. Member AIAA.

JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS

Vol. 30, No. 2, March–April 2007

419

Page 2: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

namely, solar sail and minimagnetospheric plasma propulsion, thatcan generate thrust-induced forces varying as the inverse square ofthe heliocentric distance. For both propulsion types, however, thecurrent and near term technology can guarantee only small � values.Assuming for instance �max � 0:5 (which is well beyond the currenttechnology), the maximum orbit radius attainable using an inversesquare radial thrust is only 2r0. This result substantially limits thepractical applicability of such a transfer technique, the more so forescape orbits.

For these reasons we investigate a hybrid low-thrust propulsion(HLTP) system in which a system capable of providing an inversesquare radial thrust is coupled with a SEP system. The idea ofcombining different low-thrust propulsion systems has beenintroduced by Leipold and Götz [7]. In their preliminary study [7],the HLTP concept was used to assess the feasibility of the idea androughly estimate the performance enhancements, in terms ofmissiontime, of a solar sail combined with a SEP system with respect to apure solar sail.

The primary purpose of this paper is to systematically quantify theimpact of the control parameter � on the propellant consumption for atransfer circle-to-circle problem. To make a meaningful comparisonbetween a HLTP and a conventional SEP system, these two optionsare compared in terms of payload mass fraction deliverable for agiven mission. Also, the tradeoff between payload size and trip timeis established.

Quasi-Hohmann Transfer

To better appreciate the potentialities of a hybrid propulsionsystem, we begin our discussion by considering a biimpulsivetransfer between circular coplanar orbits for a spacecraft with anauxiliary inverse square radial thrust, capable of providing a constantvalue � � �max < 1. In particular, we investigate the missionperformance, in terms of required �V and transfer time tf, for anelliptical transfer orbit tangent to both the initial and final circularorbits with radii r0 and rf, respectively. This analysis extends thefamiliar Hohmann results to a quasi-Hohmann transfer, charac-terized by a spacecraft subject to a gravitational attraction equal to�1 � �max� times that it would experience without the presence of theauxiliary propulsive system.

Introducing the ratio x≜ rf=r0 between the two orbital radii, theenergy equation yields

�V����������������=r0

p � sgn�x � 1���x � 1�

������������������������2�1� �max�x�x� 1�

s� 1���

xp � 1

�(1)

where sgn��� is the signum function. The flight time divided by theparking orbit period is equal to one-half the transfer orbit period andis given by [6]

tf

2��������������r30=��

p �����������������������������1� x�3

32�1 � �max�

s(2)

Figure 1 shows that the�V saving with respect to a case without anauxiliary propulsion system (�max � 0) is an increasing function of�max. Correspondingly, the mission time tends to increase as �max

grows large. Figure 2 specializes the preceding results for an Earth–Venus and Earth–Mars rendezvous mission.

Problem Statement

Assume that the spacecraft has a HLTP system, constituted by aprimary SEP system and an auxiliary system capable of generating amodulated purely radial thrust (that is, parallel to the sun–spacecraftdirection) and variable according to the inverse square law with thedistance r from the sun.At a generic time instant the thrust by the SEPand the auxiliary system are referred to as Te and Ta, respectively.While Te is constant, the auxiliary system produces a thrust in theform

Ta �m���r2

(3)

wherem is the (time variable) spacecraftmass, and� 2 0; �max is theratio of the auxiliary thrust force to the solar gravitational force. Themain effect of the auxiliary system, which, by assumption, has a nullpropellant consumption, is to reduce the gravitational acceleration bya factor of �1� ��.

max

max

max

0

0.050.15

0.2

0

0.150.2

0.10.05

0.1

0

0.2

Fig. 1 �V and mission time required for a quasi-Hohmann transfer.

ηmax

V

õ/r

0

Ea rth-Mars

Earth-Venus

ηmax

t f2

π√r

3 0/

Earth-Mars

Earth-Venus

µ

Fig. 2 �V variation and flight time for a quasi-Hohmann transfertowards Venus and Mars.

420 MENGALI AND QUARTA

Page 3: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

The heliocentric equations of motion for a spacecraft in a polarinertial frame T ��r; �� are

_r� u (4)

_�� v

r(5)

_u� v2

r� �1 � ����

r2� �Te

mcos� (6)

_v�� uv

r� �Te

msin� (7)

_m�� �Te

ve(8)

where � ismeasured counterclockwise from some reference position,� � �0; 1� is the electric thruster switching parameter, and � 20; 2 is the electric thruster direction angle (in the orbital plane)measured counterclockwise from the radial direction. Clearly, thesystem of differential equations (4–8) is similar to that describing thetwo-dimensionalmotion of a spacecraft equippedwith a SEP system,and widely studied in the literature (see, e.g., [8,9]), with the additionof the further control parameter � that accounts for the presence of theauxiliary propulsive system.

Our aim is to evaluate the impact of the auxiliary system on thespacecraft trajectories for a rendezvous mission between helio-centric, circular, and coplanar orbits of given radii r0 (starting orbit)and rf (target orbit). To establish a meaningful comparison, weinvestigate optimal trajectories in terms of minimum propellantconsumption necessary to accomplish the mission. From amathematical standpoint the problem is that of finding the optimalcontrol law u�t� (where t 2 t0; tf) that minimizes the propellantmass necessary to transfer the spacecraft from the initial to the finalorbit. Equivalently, the performance index

J �mf (9)

must bemaximized, wheremf is the spacecraft mass at the final time.

Optimal Steering Law

From Eqs. (4–8), the Hamiltonian associated with the problem is

H � �ru� ��

v

r� �u

�v2

r� �1 � ����

r2� �Te

mcos�

� �v

�� uv

r� �Te

msin�

�� �m

�Te

ve(10)

where �r, ��, �u, �v, and �m are the adjoint variables associated withthe state variables r, �, u, v, andm, respectively. The time derivativesof the adjoint variables are provided by the Euler–Lagrangeequations

_� � � 0 (11)

_� r ���v

r2� �u

�v2

r2� 2�1� ����

r3

�� �v

uv

r2(12)

_� u ���r � �v

v

r(13)

_� v ����

r� 2

�uv

r� �vu

r(14)

_� m � �Te

m2��u cos�� �v sin�� (15)

Assuming that the final time is left free, the boundary conditions forthe differential problem described by the equations of motion (4–8)and the Euler–Lagrange Eqs. (12–15) are initial boundary conditions

r�t0� � r0; ��t0� � u�t0� � 0 (16)

v�t0� �����������������=r0

p; m�t0� �m0 (17)

and final boundary conditions

r�tf� � rf; ���tf� � u�tf� � 0 (18)

v�tf� �����������������=rf

q; �m�tf� � 1 (19)

The preceding conditions are representative of a spacecraft leavingthe sphere of influence with zero hyperbolic excess velocity. FromEqs. (11) and (18) it follows that �� � 0.

Invoking the Pontryagin’s maximum principle, the optimalcontrol law is obtained by maximizing that portion H0 of the

Hamiltonian that explicitly depends on the control vector u≜�; �; �

H0 ≜ �u

��� � 1���

r2� �Te

mcos�

�� �v

�Te

msin� � �m

�Te

ve(20)

As far as � is concerned, the optimal control law is well known [8,9]and is given by

cos�� �u������������������2u � �2

v

p ; sin�� �v������������������2u � �2

v

p (21)

As a result, the optimal thrust vector direction provided by the SEPsystem coincides with that of the primer vector [8] whose radial andtransverse components are �u and �v, respectively. The optimalcontrol law for � is found by observing thatH0 depends linearly on �.As a result, a bang-bang control is optimal [9]

� �

8>>><>>>:1 if

������������������2u � �2

v

pm

� �m

ve> 0

0 if

������������������2u � �2

v

pm

� �m

ve� 0

(22)

A similar situation occurs with �, whose optimal control law issimply given by

����max if �u > 0

0 if �u � 0(23)

Numerical Simulations

A set of canonical units [11] have been used in the integration ofthe differential equations to reduce their numerical sensitivity. Thedifferential Eqs. (4–8) and (12–15) have been integrated in doubleprecision using a Runge–Kutta-fifth-order schemewith absolute andrelative errors of 10�12. The boundary value problem associated withthe variational problem has been solved by means of a hybridtechnique that combines gradient-based and direct methods [12].

The control laws described earlier have been applied to study therendezvous trajectories towardsVenus (rf � r♀ � 0:72333199 AU)andMars (rf � r♂ � 1:52366231 AU) for values of �max ranging in

the interval [0, 0.5]. The characteristics of the SEP system and thelaunch mass are consistent with that of the European SMART-1spacecraft, and the corresponding data have been taken from [13]. In

MENGALI AND QUARTA 421

Page 4: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

particular, a SEP thrust Te � 73 mN, an exhaust velocityve � 16:4 km=s, and an initial spacecraft mass m0 � 350 kg havebeen assumed.

To better highlight the role of the auxiliary system on the massnecessary to accomplish the mission, the following two cases havebeen considered.

Case a: The parameter � is held fixed for thewhole trajectory usinga value coincident with its maximum attainable value � � �max�const. This case is representative of a situation in which Ta cannot bemodulated nor zeroed. The optimal control law for the other twocontrol variables � and � is given by Eqs. (21) and (22).

Case b: the parameter � 2 0; �max is actually a control variableand its steering law is provided by Eq. (23).

Figures 3 and 4 show the simulation results, for the tworendezvous test missions, as a function of different values of �max. Tobetter quantify the impact of the auxiliary system over the missionperformance, the two figures also display the case corresponding tothe employment of a SEP thruster with no auxiliary system (Ta � 0)during the whole mission.

As expected, the performance in terms of final mass and missiontime are much worse in case a than in case b (in fact, the lattercoincides with the optimal condition). There are other points worthnoting. First, the performance in case a is alsoworse (except for smallvalues of �max) than that corresponding to a pure SEP thruster.Moreover, the performance in case a tends to degrade quickly as �max

is increased. This is a counterintuitive behavior because one mightexpect that a decrease in the solar gravitational attraction (as happenswhen the auxiliary system is turned on) could increase theeffectiveness of the SEP system, at least as long as missions towardsouter planets are considered. However, this is a wrong conclusion. Infact, the presence of a significant and constant value of � forces thespacecraft into an initial highly elliptic trajectory [6], setting in thisway the aphelion well beyond the radius of the target orbit, see Fig. 5(continuous lines represent case a). As a result, the electric thruster inthe initial mission phase is not used to accelerate but rather todecelerate the spacecraft, so as to counteract the negative effect of theauxiliary system. This characteristic is clearly seen in Fig. 6 in whichthe time histories of �, corresponding to case a, are shown for an

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8m

f/m

0

case (a)

case (b)

ηmax

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5200

300

400

500

600

700

800

t f[d

ays]

case (a)

case (b)

pure electric propulsion

pure electric propulsion

ηmax

Fig. 3 Results for an Earth–Venus circle-to-circle rendezvous mission.

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

ηmax

mf

/m0

case (a)

case (b)

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5200

300

400

500

600

700

800

ηmax

t f[d

ays] case (a)

case (b)

pure electric propulsion

pure electric propulsion

Fig. 4 Results for an Earth–Mars circle-to-circle rendezvous mission.

max=0.2 =0.3max

=0.4max =0.5max

Fig. 5 Earth–Mars trajectories inclusive of the start (circle) and arrival(square) spacecraft positions.

[d

eg]

α

-400

-350

-300

-250

-200

-150

-100

-50

0

50

100

[days]t0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 750

max 0.3η =

max 0.2η =

max 0.5η =

max 0.4η =

Fig. 6 Time histories of � for an Earth–Mars transfer (case a).

422 MENGALI AND QUARTA

Page 5: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

Earth–Marsmission using four different values of �max. For example,when �max � 0:4 and �max � 0:5, the electric thruster direction angleis���100 deg in the initial mission phase. Figure 5 also shows theballistic trajectories (dotted lines) that the spacecraft would follow

without the engagement of the SEP system. Note that when�max � 0:5, the trajectory turns into a parabola, that is an escapetrajectory. In fact, from the conservation of energy, the ratio of thespacecraft escape velocity vesc at instant t0 to the initial circular

velocity vc �����������������=r0

pis found to be

vesc����������������=r0

p �������������������������2�1� �max�

p(24)

Equation (24) shows that when �max � 0:5, vesc=vc � 1, whichcorresponds to attaining an escape condition.

Another interesting point emerging fromFigs. 3 and 4 is that case boffers, for Mars and Venus missions and for any values of �max, abetter performance than that obtainable from apure SEP system, bothin terms of propellant consumption and mission time. The mass andthe transfer time saving are shown in Fig. 7 where the termsmfe

andtfe refer to the final mass and flight time of a pure SEP based mission(�max � 0).

The mission trajectory and the time history of the control angle �for case b are shown in Fig. 8 for �max � 0 and �max � 0:5. We notethat the auxiliary system is only engaged in the early transfer phase(thick line of Fig. 8b) and that a similar behavior is found in all of theanalyzed trajectories. Observe that, as long as the auxiliary system isoperating, �� �max, as implied by Eq. (23).

Figure 7 shows that both the propellant mass and the flight-timesaving are increasing functions of �max. In particular, when �max �0:5 the mass saving is about 10 and 5% for Earth–Mars and Earth–Venus missions while the corresponding time saving is 95 and45 days, respectively. Figure 7 also shows a steep growth ofmass andtime savings for small values of �max, whereas the curve slopes tendto quickly decrease as �max is increased. This behavior implies that itwould not be a good idea to choose auxiliary systems with highvalues of �max. Anyway, the fundamental information about theeffectiveness of a HLTP system over a pure SEP system is ultimately

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

2

4

6

8

10

ηmax

mf

−m

f em

0

Earth-Mars transfer

100

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

20

40

60

80

100

ηmax

t fe

−t f

[day

s] Earth-Mars transfer

Earth-Venus transfer

Fig. 7 Mass and mission time savings for case b with respect to a pure

SEP mission.

a) Pure electric propulsion ( max=0) η

b) Case (b) ( max=0.5)

Fig. 8 Earth–Mars trajectories and time histories of the control angle �.

MENGALI AND QUARTA 423

Page 6: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

given in terms of payload mass deliverable for a given spacecraftmass. This subject is addressed in the next section.

Payload Mass Fraction

So far, we have quantified the advantages of an auxiliary system interms of �max, a parameter useful to define the performance of thesystem. However, the additional thrust system increases thespacecraft mass so that the question arises whether there exists abenefit in terms of payloadmass fraction deliverable.We need tofinda relationship between the value of �max and the mass of the auxiliarysystem. To this end, we assume the following mass breakdownmodel for the spacecraft

m0 �me �mprop �ma �mpay (25)

where me is the mass of the SEP system, mprop �mprop��max� is thepropellant mass (related to the SEP system) necessary to accomplishthe mission,ma �ma��max� is the mass of the auxiliary system, andmpay is the sum of the payload mass and the mass of the remainingsubsystems.

For a given SEP system (that is,me is fixed) and for a given valueof the initial spacecraft mass m0, the impact of the auxiliarypropulsive system on the value ofmpay can be estimated through thefunction fm � fm��max� defined as [see Eq. (25)]

fm��max�≜mpay �me

m0

� 1�mprop��max�m0

�ma��max�m0

(26)

Clearly, fm�0� represents the case in which a pure SEP system isavailable. Equation (26) allows one to quantify the relative “weight”of the auxiliary system, once a reasonable expression for the massfractionma=m0 is found. This can be done by introducing the type ofauxiliary system. Although, in principle, different systems may befound capable of generating a modulated radial thrust (variableaccording to an inverse square law with the distance from the sun) inthe following, we concentrate on solar sail-based applications.

Solar Sail Application

Assume that the auxiliary propulsive system is constituted by asolar sail whose effective surface A is kept constant. The solar sail iscapable of generating a pure radial thrust provided its reflector is keptperpendicular to the sunlight direction. Of course, the sail may beemployed more efficiently by using its capability of supplying atangential thrust component if tilted away from the radial direction.Nevertheless herewe assume a pure radial thrust to be consistentwiththe previously discussed model. Also, it should be noted that thepossible interference of the SEP exhaust plumewith the reflectormayvary the optical characteristics of the sail. This complex effect isneglected for the sake of simplicity.

We make use of the so-called nonperfect reflection model [14],which takes into account the optical coefficients of the real sail film.The thrustTa provided by a flat sail whose surface is perpendicular tothe sun–spacecraft line is given by

Ta � �am���

��r2

(27)

where �� is referred to as dimensionless sail loading, similar to thatintroduced in [14], �a � �0; 1� is the solar sail switching parameter,and � � 1 is a positive coefficient depending on the opticalcharacteristics of the sail film (�� 1 in the ideal case of specularreflection). For example, the optical coefficient for a sail with ahighly reflective aluminum-coated front side and a highly emissivechromium-coated back side is�� 0:90815 [15]. The dimensionless

sail loading is defined as �� ≜ ��=�, where �� ≜ 1:539 g=m2 is the

critical solar sail loading parameter (see [16], p. 40) and� ≜m=A is ageneralized sail loading. In fact, � is defined as a function of thewhole spacecraft mass and not of the sail mass alone (as is usuallydone): as such, � is not constant with time. Note that

�� � ��0

m0

m(28)

where

��0≜ ��=�m0=A� (29)

is the initial (constant) dimensionless sail loading. It represents atechnological parameter that quantifies the limits of the currenttechnology.

In Eq. (27), the solar sail switching parameter �a models thepossibility of zeroing the sail thrust by orienting the sail area parallelto the incident solar rays, thus obtaining �� 0. Accordingly, �a maybe thought of as a sort of solar sail activation parameter. The presenceof �a in the mathematical model guarantees the possibility ofrealizing an on/off control over � (recall that A is assumed to beconstant). On the other hand, the feasibility of a continuous variationof � between suitable limits by changing the effective sail area (assuggested in [6]) appears to be of difficult practical implementation.

Substituting Eq. (28) into Eq. (27) and comparing the latter withEq. (3), the following expression for � is found:

�� �a���0

m0

m� �a ~��0

m0

m(30)

where ~��0≜���0

is the initial dimensionless sail loading, weightedby the coefficient �. Equation (30) states that � depends on the

switching parameter �a, on the constant coefficient ~��0(that defines

the maximum performance of the auxiliary system), and on theinstantaneous mass ratio through the term m0=m. In particular, thepresence of the instantaneous mass ratio into the expression of �renders the optimal control problem substantially different from thatpreviously analyzed. In fact, although both the equations of motion(4–8) and theHamiltonian (10) are still formally valid, when Eq. (30)

is substituted into the Hamiltonian, the new control vector u≜�a; �; � nowmust be optimized. It is a simplematter to verify that theEuler–Lagrange Eqs. (11–14) remain unchanged, whereas thedifferential equation involving the mass costate �m becomes

_� m � �Te

m2��u cos�� �v sin�� � �a�u

~��0

��r2

m0

m2(31)

Equation (31) replaces Eq. (15). As a result, both the optimal controllaw for the electric thruster direction angle � and for the electricthruster switching parameter � are still given by Eqs. (21) and (22),respectively. On the other hand, because the Hamiltonian is linearwith respect to �a, the following bang-bang control law is obtainedfor the solar sail switching parameter

�a ��1 if �u > 0

0 if �u � 0(32)

The performance of aHLTP system, using a solar sail as the auxiliarypropulsion system, has been investigated for Earth–Mars and Earth–Venus rendezvous missions. The differences with respect to a pureSEP system, in terms of percentage mass ratio and mission time

savings, are shown in Fig. 9 as a function of the parameter ~��0.

Comparing the results of Fig. 9 with those obtained in the previoussection and summarized in Fig. 7, one discovers an almost exact

equivalence between the two figures (when �max and ~��0are

interchanged). This is an interesting circumstance as the results referto two distinct optimal control problems.

The explanation of such a behavior comes from the analysis of the

time history of the control variable �a. In fact, for all the values of ~��0in the range [0, 0.5], �a happens to be different from zero in the earlymission phase only. This amounts to saying that the auxiliarypropulsive system is operating for a fraction of the whole missiontime (on the order of 10–15%) when the spacecraft mass is about itsinitial valuem0. Recalling the optimal control law for � [see Eq. (23)]and bearing inmind that�, similarly to �a, is different from zero in thefirst mission phase only (see, for example, Fig. 8b), one has

424 MENGALI AND QUARTA

Page 7: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

� ≠ 0 ) �� �max; �a ≠ 0 ) �a � 1 (33)

so that from Eq. (30) one concludes that �max � ~��0�m=m0� ~��0

.In other terms, it turns out that there is a substantial equivalence

between �max and ~��0, and Fig. 7 can be used to quantify the

performance of a HLTP system with a solar sail as the auxiliarysystem.

Having expressed the mass fraction mf=m0 as a function of the

design parameter ~��0(or, equivalently, as a function of �max), it is

now possible to study the function fm defined through Eq. (26). Tothis end it is necessary to establish a suitable expression for the massfractionma=m0 associated with the auxiliary system, that is, with the

solar sail. First, introduce the sail assembly loading [17,18]

�sail ≜ma=A, i.e., the mass of the sail assembly (comprising the sailfilm and the required structure for storing, deploying and tensioningthe sail) per unit area. Then, Eq. (29) yields

ma

m0� �sail

�� ��0� �sail

��~��0

�(34)

The value of �sail depends on the technology employed to build thesail and, in particular, is closely related to the value of the density ofthe sail film. Currently, admissible values are on the order of25–30 g=m2, even if near-term andmidterm technology [18,19] willhopefully allow values of 5–10 g=m2. However, future outer solarsystem missions will require an assembly loading of the order of

1 g=m2, see [16], p. 95. Figure 10 shows the dependence of ~��0on

the assembly loading �sail as a function of the mass fraction ma=m0.The figure also shows the corresponding value of the sail

characteristic dimension (����A

p). The latter coincides with the side

length for a square sail configuration. Assuming, for example, anassembly loading of 25 g=m2 and a mass fraction equal to 0.5,

Fig. 10 shows that ~��0 0:028 (recall that this value is approxi-

mately equal to �max).Bearing in mind Eqs. (26) and (34) one has

fm �mf

m0

� �sail��

~��0

�(35)

The function fm has been drawn in Fig. 11 for some values of the sailloading. Figure 11 shows that, with the exception of very high-performance solar sails, function fm is always decreasing with

respect to ~��0. More precisely, @fm=@ ~��0

< 0 if �sail > 2 g=m2 forboth the analyzed missions. These results indicate that, for currentand near-midterm technology, the use of a solar sail as an auxiliarypropulsive system produces a decrease of the payload mass fractiondeliverable. In fact, the extra mass of the solar sail is greater than thepropellant mass that is saved with a pure SEP system (recall that thelatter is represented by the point fm�0� in Fig. 11). Nevertheless,

because an increase of ~��0makes the mission time decrease (see

Fig. 9), a HLTP system can be usefully employed in the mission

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

2

4

6

8

10

12

β̃σ 0

mf−

mf e

m0

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

20

40

60

80

100

β̃σ 0

t fe−

t f[d

ays]

100

Earth-Mars transfer

Earth-Venus transfer

Earth-Mars transfer

Earth-Venus transfer

Fig. 9 Mass andmission time savings for aHLTP systemwith solar sail

with respect to a pure SEP mission.

0.1

0.20.4

0.5

0.60.7

10 12 14 16 18 20 22 24 26 28 300

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1

10 12 14 16 18 20 22 24 26 28 300

50

71

87

100

112

123

132

142

150

158

√ A]

m[

0.10.2

0.3

0.40.5 0.6

0

1 2 3 4 5 6 7 8 9 100

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

1 2 3 4 5 6 7 8 9 100

112

158

194

224

250

274

296

316

336

354

√ A]

m[

m a / m 0

m a / m 0

0.3

0.7

2sail [g/m ]σ

2sail [g/m ]σ

∼ β σ0

∼ β σ

Fig. 10 Dependence of ~��0and of

�����

Ap

on �sail as a function of the mass

fraction.

˜ 0

f m

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

f m

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

1

1

5

5

10

10

15

15

20

2025

25

σsail

σsail

Earth-Venus

Earth-Mars

βσ

˜ 0βσ

Fig. 11 Dependence of fm [see Eq. (35)] on ~��0for some values of �sail

(g=m2).

MENGALI AND QUARTA 425

Page 8: Trajectory Design with Hybrid Low-Thrust Propulsion Systema).pdfTrajectory Design with Hybrid Low-Thrust Propulsion System Giovanni Mengali∗ and Alessandro A. Quarta† University

analysis to trade off the conflicting requirements between deliverablepayload mass fraction and mission time.

Conclusions

A novel concept of mission, based on a hybrid low-thrustpropulsion system, has been proposed and discussed. A solar electricpropulsion thruster has been coupled with an auxiliary system,capable of providing an inverse square radial thrust. The idea is that,by virtue of this additional thrust, the spacecraft “senses” a reducedgravitational solar force. As a result, the propellant expense for agiven heliocentric mission is typically decreased. To quantify thiseffect on a rigorous basis the problem has been cast into an optimalityframework. In particular, the steering law that minimizes thepropellant consumption for a circle-to-circle rendezvous problemhas been found using an indirect approach. Exploiting the size of thereduction of the gravitational acceleration as a control parameter, ithas been shown that an increase of that parameter provides not onlybetter performance in terms of propellant consumption but also interms ofmission time (which decreases as long as the value of the netgravitational acceleration decreases). To make a meaningfulcomparison between a hybrid and a conventional electric propulsionsystem, these two options have been compared in terms of payloadmass fraction deliverable for a given mission. To illustrate such acomparison, a solar sail has been used as an auxiliary system. It hasbeen shown that the advantage of a hybrid system is subordinated tothe employment of high-performance solar sails. Nevertheless, thereduction inmission times (whatever the solar sail characteristics be)of the new hybrid system makes this configuration attractive forfuture planetarymissions.Other types of low-thrust hybridization arepossible, such as solar electric power and minimagnetosphericplasma propulsion.

References

[1] Williams, S. N., and Coverstone-Carroll, V., “Benefits of Solar ElectricPropulsion for theNextGeneration of Planetary ExplorationMissions,”Journal of the Astronautical Sciences, Vol. 45, No. 2, April–June 1997,pp. 143–159.

[2] Williams, S. N., and Coverstone-Carroll, V., “Mars Missions UsingSolar Electric Propulsion,” Journal of Spacecraft and Rockets, Vol. 37,No. 1, Jan.–Feb. 2000, pp. 71–77.

[3] Brophy, J. R., “Advanced Ion Propulsion Systems forAffordableDeep-Space Missions,” Acta Astronautica, Vol. 52, Nos. 2–6, Jan.–March 2003, pp. 309–316.

[4] Dachwald, B., “Optimization of Interplanetary Solar SailcraftTrajectories Using Evolutionary Neurocontrol,” Journal of Guidance,Control, and Dynamics, Vol. 27, No. 1, 2004, pp. 66–72.

[5] Zubrin, R. M., and Andrews, D. G., “Magnetic Sails and InterplanetaryTravel,” Journal of Spacecraft and Rockets, Vol. 22, No. 2, 1991,pp. 197–203.

[6] McInnes, C. R., “Orbits in a Generalized Two-Body Problem,” Journalof Guidance, Control, and Dynamics, Vol. 26, No. 5, Sept.–Oct. 2003,pp. 743–749.

[7] Leipold, M., and Götz, M., “Hybrid Photonic/Electric Propulsion,”Kayser-Threde, GmBH, Rept. SOL4- TR-KTH-0001, Munich,Germany, Jan. 2002.

[8] Lawden, D. F., Optimal Trajectories for Space Navigation,Butterworths, London, 1963, pp. 54–68.

[9] Prussing, J. E., and Chiu, J. H., “OptimalMultiple-Impulse Time-FixedRendezvous Between Circular Orbits,” AIAA Paper 1984-2036,Aug. 1984.

[10] Bryson, A. E., and Ho, Y. C., Applied Optimal Control, Hemisphere,New York, 1975, Chap. 2, pp. 71–86.

[11] Seidelmann, K. P., Explanatory Supplement to the Astronomical

Almanac, University Science Books, Mill Valley, CA, 1992, pp. 696–697.

[12] Mengali, G., and Quarta, A. A., “Optimal Three-DimensionalInterplanetary Rendezvous Using Nonideal Solar Sail,” Journal of

Guidance, Control, and Dynamics, Vol. 28, No. 1, Jan.–Feb. 2005,pp. 173–177.

[13] Schoenmaekers, J., “Post-Launch Optimization of the SMART-1 Low-Thrust Trajectory to the Moon,” Proceedings of the 18th InternationalSymposium on Space Flight Dynamics, SP-548, ESA, 2004.

[14] Dachwald, B., Mengali, G., Quarta, A. A., and Macdonald, M.,“Parametric Model and Optimal Control of Solar Sails with OpticalDegradation,” Journal of Guidance, Control, and Dynamics (to bepublished).

[15] Wright, J. L., Space Sailing, Gordon and Breach, Philadelphia, 1992,pp. 227–233.

[16] McInnes, C. R., Solar Sailing: Technology, Dynamics and Mission

Applications, Springer–Praxis Series in Space Science andTechnology,Springer–Verlag, Berlin, 1999, pp. 40, 95.

[17] McInnes, C. R., “Payload Mass Fractions for Minimum-Time Trajectories of Flat and Compound Solar Sails,” Journal of

Guidance, Control, and Dynamics, Vol. 23, No. 6, 2000, pp. 1076–1078.

[18] Dachwald, B., “Solar Sail Performance Requirements for Missions tothe Outer Solar System and Beyond,” Proceedings of the 55th

International Astronautical Congress, International AstronauticalCongress Paper IAC-04-S.P.11, Oct. 2004.

[19] Gershman, R., and Seybold, C., “Propulsion Trades for Space ScienceMissions,” Acta Astronautica, Vol. 45, Nos. 4–9, 1999, pp. 541–548.

426 MENGALI AND QUARTA