<<

MULTIPLE REVOLUTION LAMBERT´S TARGETING PROBLEM: AN ANALYTICAL APPROXIMATION

Claudio Bombardelli, Juan Luis Gonzalo, and Javier Roa

Space Dynamics Group, School of Aeronautical Engineering, Technical University of Madrid

ABSTRACT its initial guess. A quite detailed historical review on the dif- Solving a Lambert’s Targeting Problem (LTP) consists of ob- ferent contributions to the method can be found in the intro- taining the minimum required single-impulse delta-V to mod- duction of reference[1]. ify the trajectory of a spacecraft in order to transfer it to a se- When focusing on the design and optimization of space lected orbital position in a fixed time of flight. We present an trajectories one encounters a subclass of Lambert’s Problem approximate analytical method for rapidly solving a generic referred to as Lambert’s Targeting Problem (LTP). It consists LTP with limited loss of accuracy. The solution is built on an of obtaining the minimum required single-impulse delta-V to optimum single-impulse time-free transfer with a D-matrix modify the trajectory of a spacecraft in order to transfer it to phasing correction. Because it exhibits good accuracy (both a selected orbital position in a fixed time of flight. If the so- in terms of delta-V and time of flight) near locally optimum lutions of a Lambert’s Problem have been obtained the LTP transfer conditions the method is useful for rapidly obtaining is solved right away by picking the one providing minimum low delta-V solutions for interplanetary trajectory optimiza- delta-V magnitude within the maximum number of revolu- tion. In addition, the method can be employed to provide a tions physically allowed by the transfer time constraint. The first guess solution for enhancing the convergence of reason for distinguishing between LP and LTP is mostly re- an accurate numerical Lambert solver. lated to the method proposed in the present work, which is Index Terms— Lambert’s Problem, D Matrix, Quartic concerned with the solution of an LTP but not the general so- Equation, Targeting lution of an LP. Solving an LTP with the highest possible computational 1. INTRODUCTION efficiency is key to the design of optimized interplanetary tra- jectories, and, in particular, the ones including multiple grav- Lambert’s problem (LP), i.e. the determination of the fi- ity assists and deep space maneuvers (DSMs). For these types nite set of linking two position vectors in a two-body of problems, the complete sampling (grid search) of the multi- gravitational field with a specified transfer time, is of funda- dimensional solution space implies the solution of a num- mental importance in and has been studied ber of LTPs often exceeding the capability of even the for more than two centuries. If one accounts for multiple advanced computational means available today. One exam- revolutions and both prograde and retrograde transfers Lam- ple is the design and optimization of multiple D R  bert’s problem provides 2 Nub + Nub + 1 solutions where trajectories with multiple DSMs like the Messenger trajec- D R Nub,Nub are, respectively, the maximum number of revo- tory to , which employed seven trajectory arcs with lutions that a direct (D) and retrograde (R) orbits can have five deterministic DSM with delta-V larger than 70 m/s[2]. before reaching the target location[1]. That means that if The sheer number of LTPs to be solved in this case makes long-duration transfer arcs are sought the problem becomes the problem computationally intractable if a brute-force, grid- more difficult to solve, the longer so the transfer time consid- sampling approach is to be followed (see Conway’s book[3], ered. page 202). Another relevant application for a high-efficiency The solutions of a Lambert’s problem are to be obtained LTP solver is the design of “inter-satellitary” trajectories at numerically and are essential for various tasks in astrodynam- the giant planets where multi-revolutions transfers are com- ics, including the design of interplanetary trajectories. Most mon and the number of possible flyby sequences make the efforts in the literature have been devoted to the improvement design and optimization task “computationally gigantic”[4]. of the convergence speed of its numerical solution using dif- It is clear that even a modest increase of efficiency in an LTP ferent choices for the iteration variable and ways to improve solution algorithm would be extremely beneficial when fac- ing these computational challenges. In addition, any intrinsic Eq. (2) is a quartic algebraic equation that can be solved pruning capability that the algorithm could offer in order to analytically using different methods (e.g. by Lagrange resol- discard unintresting transfer arcs a priory is highly welcome. vent) or numerically (e.g. with Newton’s or Halley’s method). Recent advances towards a highly efficient LTP solver Once solved, the optimal chordal and radial components of v1 have been made by these authors[5], who proposed an ap- are determined using Eq. (3) and the following relation: proximate analytical solution of the LTP based on Battin’s 1 r2r r ∆θ optimum single-impulse transfer solution and a linear phas- = 1 2 x cos , (4) ing correction through the use of a novel Keplerian error state vc1 µc 2 transition matrix[6],[7]. The solution provides a useful tool Having determined the optimum v1 the total single- for preliminary trajectory design when high accuracy is not impulse transfer velocity vector is obtained as: a must as long as the main features of the solution are pre- served. ∆vB = v1 − v0, (5) We review the steps leading to the derivation of the analyt- ical LTP solver providing additional formulas and examples while the angular momentum and of the not previously described. transfer are obtained as:

2. ANALYTICAL LTP SOLVER h = huh = r1 × v1

Let us consider the Lambert’s Targeting Problem (LTP) of a v1 × h e = eue = − ur1. spacecraft departing from a point r1 with initial velocity v0 µ and arriving at a point r in a fixed time of flight ∆t while 2 L The true and eccentric anomalies at r and r can now be moving in a Keplerian gravitational field of gravitational pa- 1 2 determined from the relations: rameter µ. The required delta-V targeting impulse ∆vL can be broken down into an optimum time-free impulsive delta-V (∆vB) and a phsing correction term (∆vC ): cos θ1,2 = ur1,2 · ue; sin θ1,2 = (ue × ur1,2) · uh,

∆vL = ∆vB + ∆vC . (1) √ 2 e − cos θ1,2 1 − e sin θ1,2 Time-free delta-V cos E1,2 = ; sin E1,2 = . 1 + e cos θ1,2 1 + e cos θ1,2 The optimum time-free impulsive delta-V (∆vB) can be ob- tained analytically following the method proposed by Battin, The semi-major axis and are readily com- which reduces to the solution of the quartic equation[8]: puted as: r x4 − P x3 + Qx − 1 = 0 (2) r1 µ a = , n = 3 1 − e cos E1 a where x > 0 is the square of the ratio between the skew- radial and skew-chordal components of the optimum depar- Finally the transfer time satisfies Kepler’s equation: ture velocity E2 − E1 − e (sin E2 − sin E1) v ∆tB = . x2 = ρ1 , (3) n vc1 For hyperbolic orbits (e > 1), all preceding relations are while: still valid after considering the relation between hyperbolic and elliptic : r 2r1r2 ∆θ P = cos (v0π · ur1) , µc 2 H = iE. (6) r2r r ∆θ Q = 1 2 cos (v · u ) , µc 2 0π c D matrix targeting

with r1, r2, ∆θ, c indicating, respectively, the magnitudes Well-designed transfer arcs are characterized by high phasing of the departure and arrival location, the angular width of the efficiencies or, equivalently, small magnitude velocity cor- transfer arc and its chordal length, while v0π, ur1, uc are, re- rections ∆vC . When this is the case one can obtain a suf- spectively, the transfer plane component of the spacecraft ini- ficiently accurate estimation of ∆vC starting from the opti- tial velocity v0, the radial unit vector at departure and the mum single-impulse transfer trajectory. The sought maneu- chordal unit vector from r1 to r2. ver correction will be the one adjusting the trajectory phasing without changing the orbit radius at the arrival where: θ2. The phasing correction to be applied is best estimated as: ∆E = E − E1 δt = ∆tL − ∆tB − NTB, (7) ∆SE = sin E − sin E1 where N is the total number of full revolutions of the clos- est optimum single-impulse transfer: ∆CE = cos E − cos E1   ∆tL − ∆tB N = , ∆S2E = sin 2E − sin 2E1 TB with [] denoting the nearest integer function. ∆C2E = cos 2E − cos 2E1. The velocity correction ∆vC associated with δt can be obtained from the linear relation linking the radial position shift (δr) and time delay (δt) of a Keplerian orbit at true The necessary targeting delta-V conditions are obtained anomaly θ to the applied delta-V maneuver along the radial by inversion of Eq.(8) after setting zero radial error et en- counters and the required targeting time delay in Eq.(7). In (δvr,1) and transversal direction (δvθ,1) at true anomaly θ1 [6],[7]: this manner:         δr (θ) δvr,1 δvr,1 ∗ 0 ≈ D (θ, θ1) . (8) = D . δt (θ) δvθ,1 δvθ,1 δt The D matrix is essentially a Keplerian error state tran- where we have introduced the “guidance matrix” D∗ = D−1. sition matrix where time (t) is a state variable together with Finally, the sought velocity correction yields: the radial distance (r) and the radial and transversal velocity (v , v ) while the independent variable is the angular position r θ ∆vC ≈ δvr,1ur1 + δvθ,1uθ. (9) θ on the reference orbit (conveniently measured from the ref- erence orbit eccentricity vector). Eq(9), together with (5), provide the approximate solution The D matrix terms were first obtained in reference[6] and of the LTP (Eq.(1)). are here rewritten in a more compact way in terms of angular momentum, radial distance and classical of the reference orbit: Performance of the method r2 D = sin (θ − θ ) , 1,1 h 1 The performance of the method, in terms of accuracy and speed, has been tested by producing pork-chop plots of 2 r r1 2 − 2 cos (θ − θ1) − e sin θ1 sin (θ − θ1) direct transfer trajectories to and to the asteroid 65803 D = × . 1,2 h a (1 − e2) Didymos[9]. In all cases, contour plots were produced for both the departure energy (C3) and arrival excess velocity at the target celestial body (∆V ). The plots were obtained by a4 1-e2 a 1000×1000 discretization of the departure × time D2,1 = [− (4∆CE − e∆C2E) (cos E1 − e) of flight coordinates leading to a one million Lambert solver 2r h2 1 calls. The analytical formulas derived in this article were 2   + 6e∆E − 4 1 + e ∆SE + e∆S2E sinE1 , coded in fortran using a GNU gfortran compiler (version 5.3.1) in simple precision with an Intel Core processor i7- [email protected]. The use of simple precision fits well within √ the scope of a preliminary optimization tool and its future 4 2 a 1 − e  2 2 GPU implementation. D2,2 = 2 12 1 − e ∆E − 3e ∆S2E 4r1h In order to assess the accuracy of the method and its com- 3 2  +6e ∆SE + 2 2 − e sinE1 − esin2E1 × putational speed three very fast Lambert solver algorithms were considered: the classical Gooding’s method[10], the (4∆CE − e∆C2E ) + (4 cos E1 − e cos 2E1) × recently published Izzo’s method[11] and the Arora-Russell method[1]. The double-precision fortran versions of the 2  e∆S2E − 2 ∆SE 2 − e , two former methods were kindly made available by Jacob 1000 Table 1. Simplified ephemerides for the considered celestial bodies 14 900 C =7.8344 km2/s2 Mars 3 13 800 (t 05−08−2026) dep reference epoch (MJD) 58849 58849 57200 700 12

semi-major axis (AU) 1.00 1.52 1.64 600 11 C =8.8803 km2/s2 eccentricity 0.0167 0.0934 0.384 500 3 (t 11−27−2028) 2 2 dep 400 C =9.1575 km /s 10 inclination (deg) 0.00280 1.85 3.41 transfer time [days] 3 (t 10−27−2026) 2 2 dep C =8.9610 km /s 300 3 (t 12−05−2028) 9 argument of pericenter (deg) 287 285 319 dep 200 longitude of ascending node (deg) 176 49.5 73.2 8 100 2025.5 2026 2026.5 2027 2027.5 2028 2028.5 2029 2029.5 2030 (deg) 357 247 190 departure date [year]

1000 14

900 C =7.8344 km2/s2 Williams through the Fortran Astrodynamics Toolkit git- 3 13 800 (t 05−08−2026) hub repository[12]. A double-precision fortran code for the dep 700 12

Arora-Russell method was kindly provided to the authors by 600 11 C =8.8803 km2/s2 Nitin Arora and Ryan Russell. 500 3 (t 11−27−2028) dep C =9.1599 km2/s2 Simplified ephemerides, reported in Table 1, were used in 400 3 10 transfer time [days] (t 10−27−2026) 2 2 dep C =8.9610 km /s all cases in order to ease the reproducibility of the results. 300 3 (t 12−05−2028) 9 dep It is important to underline that in order to obtain suf- 200 8 100 ficiently accurate approximations the analytical LTP solver 2025.5 2026 2026.5 2027 2027.5 2028 2028.5 2029 2029.5 2030 needs to be run starting from two different optimum single departure date [year] impulse problems. For the departure C3 estimation problem the targeting is clearly done over an optimum single impulse Fig. 1. Numerical (top) and analytical (bottom) pork-chop problem of a trajectory starting from Earth. Conversely, for plot for the injection energy (C3) of an Earth-Mars transfer. the arrival excess velocity estimation problem a reversed tra-

1000 jectory starting from the target body and headed towards the ∆ V=2.8267 km/s ∆V=2.3764 km/s 4.2 900 (t 04−14−2026) dep (t 07−22−2028) Earth is employed (meaning that the sign of all velocity vector dep ∆V=2.3323 km/s 4 800 (t 09−30−2028) components has to be switched). dep 3.8 700 ∆ ∆V=3.5716 km/s V=2.7444 km/s 3.6 (t 12−16−2025) (t 04−22−2026) 600 dep dep 3.4

500 3.2 ∆V=2.9493 km/s (t 08−29−2028) 400 ∆V=2.5592 km/s dep 3 Mars trajectories transfer time [days] (t 07−15−2026) 300 dep 2.8 Figures 1 and 2 show the analytically and numerically com- 200 2.6 2.4 100 puted pork-chop plots for an Earth-Mars transfer in the years 2025.5 2026 2026.5 2027 2027.5 2028 2028.5 2029 2029.5 2030 2025-2030. Contour plots are given for both departure energy departure date [year]

1000 (fig. 1) and arrival excess velocity at Mars (2). The quality ∆ V=2.9462 km/s ∆V=2.3764 km/s 4.2 900 (t 04−14−2026) dep (t 07−22−2028) of the analytical pork-chop plot is remarkable and all main dep ∆V=2.3324 km/s 4 800 (t 09−30−2028) features appear to be well reproduced. In addition the delta-V dep 3.8 700 ∆ ∆V=3.5650 km/s V=2.7444 km/s 3.6 (t 04−20−2026) and launch date error for all main local optimum solutions are (t 12−14−2025) dep 600 dep 3.4 negligible. 500 ∆V=2.9465 km/s 3.2 (t 08−28−2028) As for the computing time the method, provides a reduc- 400 ∆V=2.5588 km/s dep 3 transfer time [days] (t 07−15−2026) dep tion of a factor of 4.3, 5.1, 10.0 with respect to the method 300 2.8 of Arora-Russell, Izzo and Gooding, respectively. The maxi- 200 2.6 2.4 100 mum number of revolutions for the previous methods (which 2025.5 2026 2026.5 2027 2027.5 2028 2028.5 2029 2029.5 2030 mildly affects Arora-Russell and Izzo’s method but does af- departure date [year] fect Gooding’s one considerably) was set to 2 in this case. Fig. 2. Numerical (top) and analytical (bottom) pork-chop plot for the arrival excess velocity (∆V ) of an Earth-Mars Didymos trajectories transfer. Figures 3 and 4 show the analytically and numerically com- puted pork-chop plots for a transfer to the binary asteroid although the accuracy of the delta-V and launch date near 65803 Didymos in the years 2019-2022. Contour plots are minimum condition is not as remarkable as in the Mars case, given for both departure energy (fig. 3) and arrival excess ve- especially for the C3 plot. This is due to the low phasing effi- locity (4). Similarly to the previous case the structure of the ciency (as low as ∼67%) of some of the best solutions found. pork-chop plots is well reproduced by the analytical method In any case, the error remains within the limit of a few per- 900 4.0, 5.3, 11.1 with respect to the method of Arora-Russell, 35 800 2 2 Izzo and Gooding, respectively. The difference when com- C =1.8545 km /s 2 2 3 C =2.1082 km /s (t 01−22−2021) 30 700 3 dep pared to the Mars case is due to the problem dependency of (t 08−16−2020) dep 600 25 the numerical methods (the computational time of the ana- C =24.948 km2/s2 3 lytical method is virtually constant). For problems involving 500 (t 10−28−2020) 20 dep multi-rev solutions with a high number of revolutions the dif- 400 15 transfer time [days] C =3.1844 km2/s2 ference in performance may be substantial. 300 3 (t 08−26−2021) 10 dep 200 5

100 2019.5 2020 2020.5 2021 2021.5 2022 3. CONCLUSIONS departure date [year]

900 A novel approximate analytical solution of the Lambert’s Tar- 35 800 C =1.7534 km2/s2 geting Problem (LTP) was presented and investigated. Al- 3 C =1.9623 km2/s2 3 (t 01−24−2021) 30 700 dep though not yet extensively tested, the solution appears to be (t 08−16−2020) dep 600 25 able to reproduce the most interesting parts of a generic pork-

500 20 chop plot for interplanetary transfer trajectories. In particular, C =25.150 km2/s2 3 (t 10−27−2020) 400 dep 15 it offers remarkably good accuracy (considering its analyti- transfer time [days] 300 C =2.8281 km2/s2 3 10 cal character) in both estimated delta-V and launch date at (t 08−25−2021) dep 200 local minimum points. As for computational speed, the pro- 5

100 posed solution is at least 4 times faster than the most efficient 2019.5 2020 2020.5 2021 2021.5 2022 departure date [year] Lambert solvers currently available and more than one order of magnitude faster than Gooding’s method. A limited effort Fig. 3. Numerical (top) and analytical (bottom) pork-chop has been done in order to increase the efficiency of its imple- plot for the injection energy (C3) of an Earth-Didymos trans- mentation, which opens interesting possibilities for the future. fer. The limited accuracy of the method, which is obviously its main drawback, should be traded off with the need of compu-

900 tationally effective solutions when dealing with massive and 4.5 800 highly complex optimization problems. Many of these as- ∆V=0.6552 km/s 4 700 (t 11−21−2020) pects will need to be studied extensively in the future. dep 3.5 600 ∆V=0.6552 km/s (t 11−27−2020) 3 dep 500 2.5 Acknowledgements 400 2 transfer time [days] 300 1.5 This work has been supported by the Spanish Ministry of 200 1 Economy and Competitiveness within the framework of the 100 2019.5 2020 2020.5 2021 2021.5 2022 departure date [year] research project “Dynamical Analysis, Advanced Orbital Propagation, and Simulation of Complex Space Systems” 900 4.5 (ESP2013-41634-P). The authors also want to thank the 800 4 ∆V=0.6815 km/s funding received from the European Union Seventh Frame- 700 (t 11−21−2020) dep 3.5 work Programme (FP7/2007-2013) under grant agreement 600 ∆V=0.6544 km/s (t 11−27−2020) 3 dep N 607457 (LEOSWEEP). Last but not least, we thank Ryan 500 2.5 Russell (University of Texas at Austin), Nitin Arora (Jet 400 2

transfer time [days] Propulsion Laboratory) and Jacob Williams (NASA Johnson) 300 1.5 for kindly providing fortran implementations of different 200 1 Lambert solvers. 100 2019.5 2020 2020.5 2021 2021.5 2022 departure date [year] 4. REFERENCES Fig. 4. Numerical (top) and analytical (bottom) pork-chop plot for the arrival excess velocity (∆V ) of an Earth-Didymos [1] Nitin Arora and Ryan P Russell, “A fast and robust mul- transfer. tiple revolution lambert algorithm using a cosine trans- formation,” Paper AAS, pp. 13–728, 2013. 1, 2 cent, which may still be acceptable at a preliminary mission [2] Andrew G Santo, Robert E Gold, Ralph L Mc- design stage. Nutt, Sean C Solomon, Carl J Ercol, Robert W Far- The computational speed reduction factor in this case is quhar, Theodore J Hartka, Jason E Jenkins, James V McAdams, Larry E Mosher, et al., “The messenger mis- sion to mercury: Spacecraft and mission design,” Plane- tary and Space Science, vol. 49, no. 14, pp. 1481–1500, 2001. 1 [3] Bruce Conway, Spacecraft trajectory optimization, vol. 29, Cambridge University Press, 2010. 1 [4] Andrew F Heaton, Nathan J Strange, James M Longuski, and Eugene P Bonfiglio, “Automated design of the europa orbiter tour,” Journal of Spacecraft and Rockets, vol. 39, no. 1, pp. 17–22, 2002. 1

[5] Claudio Bombardelli, Javier Roa, and Juan Luis Gon- zalo, “Approximate analytical solution of the multi- ple revolution lambert’s targeting problem,” in 26th AAS/AIAA Space Flight Mechanics Meeting, Napa, CA, 2016. 1

[6] Claudio Bombardelli, “Analytical formulation of im- pulsive collision avoidance dynamics,” Celestial Me- chanics and Dynamical Astronomy, vol. 118, no. 2, pp. 99–114, 2014. 1, 2, 2 [7] Claudio Bombardelli and Javier Hernando-Ayuso, “Op- timal impulsive collision avoidance in ,” Journal of Guidance, Control, and Dynamics, vol. 38, no. 2, pp. 217–225, 2015. 1, 2 [8] Richard H. Battin, An introduction to the mathematics and methods of astrodynamics, Aiaa, 1999. 2

[9] A Galvez, I Carnelli, P Michel, AF Cheng, C Reed, S Ulamec, J Biele, P Zbrll, and R Landis, “Aida: The as- teroid impact & deflection assessment mission,” in Eu- ropean Planetary Science Congress, EPSC2013-1043, 2013. 2

[10] R.H Gooding, “A procedure for the solution of lambert’s orbital boundary-value problem,” and Dynamical Astronomy, vol. 48, no. 2, pp. 145–165, 1990. 2 [11] Dario Izzo, “Revisiting lambert’s problem,” Celestial Mechanics and Dynamical Astronomy, vol. 121, no. 1, pp. 1–15, 2014. 2 [12] Jacob Williams, “Fortran-astrodynamics-toolkit,” https://github.com/jacobwilliams/ Fortran-Astrodynamics-Toolkit, Accessed: 2016-02-02. 2