Abstract

Numerical and first-order analytical results are presented for optimal low-thrust limited-power trajectories in a gravity field that includes the second zonal harmonic 𝐽2 in the gravitational potential. Only transfers between orbits with small eccentricities are considered. The optimization problem is formulated as a Mayer problem of optimal control with Cartesian elements—position and velocity vectors—as state variables. After applying the Pontryagin Maximum Principle, successive canonical transformations are performed and a suitable set of orbital elements is introduced. Hori method—a perturbation technique based on Lie series—is applied in solving the canonical system of differential equations that governs the optimal trajectories. First-order analytical solutions are presented for transfers between close orbits, and a numerical solution is obtained for transfers between arbitrary orbits by solving the two-point boundary value problem described by averaged maximum Hamiltonian, expressed in nonsingular elements, through a shooting method. A comparison between analytical and numerical results is presented for some maneuvers.

1. Introduction

This paper presents numerical and analytical results for optimal low-thrust limited-power transfers between orbits with small eccentricities in a noncentral gravity field that includes the second zonal harmonic 𝐽2 in the gravitational potential. This study has been motivated by the renewed interest in the use of low-thrust propulsion systems in space missions verified in the last two decades due to the development and the successes of two space mission powered by ionic propulsion: Deep Space One and SMART 1. Several researchers have obtained numerical and sometimes analytical solutions for a number of specific initial orbits and specific thrust profiles in central or noncentral gravity field [126].

It is assumed that the thrust direction is free and the thrust magnitude is unbounded [5, 27], that is, there exist no constraints on control variables. It is supposed that the orbital changes caused by the thrust and the perturbations due to the oblateness of the central body have the same order of magnitude. The optimization problem is formulated as a Mayer problem of optimal control with position and velocity vectors—Cartesian elements—as state variables. After applying the Pontryagin Maximum Principle [28], successive Mathieu transformations are performed and a suitable set of orbital elements is introduced. Hori method—a perturbation technique based on Lie series [29]—is applied in solving the canonical system of differential equations that governs the optimal trajectories. For transfers between close quasicircular orbits, new set of nonsingular orbital elements is introduced and the new Hamiltonian and the generating function, determined through the algorithm of Hori method, are transformed to these new elements through Mathieu transformations. First-order analytical solutions are obtained for transfers between close quasicircular orbits, considering maneuvers between near-equatorial orbits or nonequatorial orbits. These solutions are given in terms of systems of algebraic equations involving the imposed variations of the orbital elements, the duration of the maneuvers, the effects of the oblateness of the central body, and the initial values of the adjoint variables. The study of these transfers is particularly interesting because the orbits found in practice often have a small eccentricity, and the problem of slight modifications (corrections) of these orbits is frequently met [5]. For transfers between arbitrary orbits, a slightly different set of nonsingular elements is introduced, and the two-point boundary value problem described by averaged maximum Hamiltonian, expressed in nonsingular elements, is solved through a shooting method. A comparison between analytical and numerical results is presented for some maneuvers.

2. Optimal Space Trajectories

A low-thrust limited-power propulsion system, or LP system, is characterized by low-thrust acceleration level and high specific impulse [5]. The ratio between the maximum thrust acceleration and the gravity acceleration on the ground, 𝛾max/𝑔0, is between 104 and 102. For such system, the fuel consumption is described by the variable 𝐽 defined as 1𝐽=2𝑡𝑡0𝛾2𝑑𝑡,(2.1) where γ is the magnitude of the thrust acceleration vector 𝜸, used as control variable. The consumption variable 𝐽 is a monotonic decreasing function of the mass 𝑚 of space vehicle: 𝐽=𝑃max1𝑚1𝑚0,(2.2) where 𝑃max is the maximum power, and 𝑚0 is the initial mass. The minimization of the final value of the fuel consumption 𝐽𝑓 is equivalent to the maximization of 𝑚𝑓.

The general optimization problem concerned with low-thrust limited-power transfers (no rendezvous) will be formulated as a Mayer problem of optimal control by using Cartesian elements as state variables. Consider the motion of a space vehicle 𝑀 powered by a limited-power engine in a general gravity field. At time 𝑡, the state of the vehicle is defined by the position vector 𝐫(𝑡), the velocity vector 𝐯(𝑡), and the consumption variable 𝐽. The geometry of the transfer problem is illustrated in Figure 1. The control 𝜸 is unconstrained, that is, the thrust direction is free, and the thrust magnitude is unbounded.

The optimization problem is formulated as follows: it is proposed to transfer the space vehicle 𝑀 from the initial state (𝐫0,𝐯0,0) at the initial time 𝑡0=0 to the final state (𝐫𝑓,𝐯𝑓,𝐽𝑓) at the specified final time 𝑡𝑓, such that the final consumption variable 𝐽𝑓 is a minimum. The state equations are 𝑑𝐫𝑑𝑡=𝐯,𝑑𝐯𝑑𝑡=𝐠(𝐫)+𝜸,𝑑𝐽=1𝑑𝑡2𝛾2,(2.3) where 𝐠(𝐫) is the gravity acceleration. For transfers between orbits with small eccentricities, the initial and final conditions will be specified in terms of singular orbital elements introduced in next sections. It is also assumed that 𝑡𝑓𝑡0 and the position of the vehicle in the initial orbit are specified.

According to the Pontryagin Maximum Principle [28], the optimal thrust acceleration 𝜸 must be selected from the admissible controls such that the Hamiltonian function 𝐻 reaches its maximum. The Hamiltonian function is formed using (2.3), 𝐻=𝐩𝑟𝐯+𝐩𝑣1(𝐠(𝐫)+𝜸)+2𝑝𝐽𝛾2,(2.4) where 𝐩𝑟,𝐩𝑣, and 𝑝𝐽 are the adjoint variables, and dot denotes the dot product. Since the optimization problem is unconstrained, 𝜸 is given by 𝜸𝐩=𝑣𝑝𝐽.(2.5) The optimal thrust acceleration 𝜸 is modulated [5], and the optimal trajectories are governed by the maximum Hamiltonian function 𝐻, obtained from (2.4) and (2.5): 𝐻=𝐩𝑟𝐯+𝐩𝑣𝑝𝐠(𝐫)𝑣22𝑝𝐽.(2.6)

The consumption variable 𝐽 is ignorable, and 𝑝𝐽 is a first integral. From the transversality conditions, 𝑝𝐽(𝑡𝑓)=1; thus, 𝑝𝐽(𝑡)=1.(2.7) Equation (2.6) reduces to 𝐻=𝐩𝑟𝐯+𝐩𝑣𝑝𝐠(𝐫)+𝑣22.(2.8) The consumption variable 𝐽 is determined by simple quadrature.

For noncentral field, the gravity acceleration is given by [30] 𝜇𝐠(𝐫)=𝑟3𝐫+𝜕𝑉2𝜕𝐫𝑇,(2.9) with the disturbing function 𝑉2 defined by 𝑉2=𝜇𝐽2𝑎𝑒2𝑟3𝑃2(sin𝜑),(2.10) where 𝜇 is the gravitational parameter, 𝑎𝑒 is the mean equatorial radius of the central body, 𝐽2 is the coefficient for the second zonal harmonic of the potential, 𝜑is the latitude, and 𝑃2 is the Legendre polynomial of order 2. (𝜕𝑉2/𝜕𝐫) is a row vector of partial derivatives, and superscript 𝑇 denotes its transpose.

Using (2.8) and (2.9), the maximum Hamiltonian function can be put in the following form: 𝐻=𝐻0+𝐻𝐽2+𝐻𝛾,(2.11) where 𝐻0=𝐩𝑟𝐯𝐩𝑣𝜇𝑟3𝐻𝐫,𝐽2=𝐩𝑣𝜕𝑉2𝜕𝐫𝑇,𝐻𝛾=𝑝𝑣22.(2.12)𝐻0 is the undisturbed Hamiltonian function, 𝐻𝐽2 and 𝐻𝛾 are disturbing functions concerning the oblateness of the central body and the optimal thrust acceleration, respectively. 𝐻𝐽2 and 𝐻𝛾 are assumed to be of the same order of magnitude.

3. Transformation from Cartesian Elements to a Set of Orbital Elements

Consider the canonical system of differential equations governed by the undisturbed Hamiltonian 𝐻0: 𝑑𝐫𝑑𝑡=𝐯,𝑑𝐯𝜇𝑑𝑡=𝑟3𝐫,𝑑𝐩𝐫=𝜇𝑑𝑡𝑟3𝐩𝑣𝐩3𝑣𝐞𝐫𝐞𝐫,𝑑𝐩𝑣𝑑𝑡=𝐩𝑟,(3.1) where 𝐞𝐫 is the unit vector pointing radially outward of the moving frame of reference (Figure 2). The general solution of the state equations is well known in Astrodynamics [31], and the general solution of the adjoint equations is obtained through properties of generalized canonical systems [32], as described in the appendix. Thus, 𝑎𝐫=1𝑒2𝐞1+𝑒cos𝑓𝐫,(3.2)𝐯=𝜇𝑎1𝑒2𝑒sin𝑓𝐞𝐫+(1+𝑒cos𝑓)𝐞𝐬,𝐩(3.3)𝐫=𝑎𝑟22𝑎𝑝𝑎+1𝑒2𝑝cos𝐸𝑒+𝑟𝑎sin𝑓𝑒𝑝𝜔1𝑒3cos𝐸1𝑒2𝑝𝑀𝐞𝐫+sin𝑓𝑎𝑝𝑒(𝑒+cos𝑓)𝑎𝑒1𝑒2𝑝𝜔+1𝑒2cos𝑓𝑝𝑎𝑒𝑀𝐞𝐬+1𝑎1𝑒2𝑎𝑟𝑝sin𝐸𝐼𝑝cos𝜔+Ωsin𝐼𝑝𝜔+cot𝐼sin𝜔1𝑒2𝑎𝑟𝑝cos𝐸𝐼𝑝sin𝜔Ωsin𝐼𝑝𝜔𝐞cot𝐼cos𝜔𝐰,𝐩(3.4)𝐯=1𝑛𝑎1𝑒22𝑎𝑒sin𝑓𝑝𝑎+1𝑒2𝑝sin𝑓𝑒1𝑒2cos𝑓𝑒𝑝𝜔+1𝑒23/2𝑒cos𝑓2𝑒𝑝1+𝑒cos𝑓𝑀𝐞𝐫+2𝑎1𝑒2𝑎𝑟𝑝𝑎+1𝑒2(cos𝑓+cos𝐸)𝑝𝑒+1𝑒2sin𝑓𝑒11+𝑝1+𝑒cos𝑓𝜔1𝑒2𝑝𝑀𝐞𝐬+𝑟𝑎cos(𝜔+𝑓)𝑝𝐼+𝑟𝑎𝑝sin(𝜔+𝑓)Ωsin𝐼𝑝𝜔𝐞cot𝐼𝐰,(3.5) where 𝐞𝐬 and 𝐞𝐰 are unit vectors along circumferential and normal directions of the moving frame of reference, respectively; 𝑎 is the semimajor axis, 𝑒 is the eccentricity, 𝐼 is the inclination of orbital plane, Ω is the longitude of the ascending node, 𝜔is the argument of pericenter, 𝑓 is the true anomaly, 𝐸 is the eccentric anomaly, 𝑀 is the mean anomaly, 𝑛=𝜇/𝑎3 is the mean motion, and (𝑟/𝑎),(𝑟/𝑎)sin𝑓, and so forth are functions of the elliptic motion which can be expressed explicitly in terms of the eccentricity and the mean anomaly through Lagrange series [31]. The true, eccentric and mean anomalies are related through the equations: 𝑓tan2=1+𝑒𝐸1𝑒tan2,𝑀=𝐸𝑒sin𝐸.(3.6) The unit vectors 𝐞𝐫,𝐞𝐬, and 𝐞𝐰 of the moving frame of reference are written in the fixed frame of reference as 𝐞𝐫+𝐞=(cosΩcos(𝜔+𝑓)sinΩsin(𝜔+𝑓)cos𝐼)𝐢(sinΩcos(𝜔+𝑓)+cosΩsin(𝜔+𝑓)cos𝐼)𝐣+sin(𝜔+𝑓)sin𝐼𝐤,𝐬𝐞=(cosΩsin(𝜔+𝑓)+sinΩcos(𝜔+𝑓)cos𝐼)𝐢+(sinΩsin(𝜔+𝑓)+cosΩcos(𝜔+𝑓)cos𝐼)𝐣+cos(𝜔+𝑓)sin𝐼𝐤,𝐰=sinΩsin𝐼𝐢cosΩsin𝐼𝐣+cos𝐼𝐤.(3.7)

Equations (3.2)–(3.5) define a Mathieu transformation between the Cartesian elements (𝐫,𝐯,𝐩𝐫,𝐩𝐯) and the orbital ones (𝑎,𝑒,𝐼,Ω,𝜔,𝑀,𝑝𝑎,𝑝𝑒,𝑝𝐼,𝑝Ω,𝑝𝜔,𝑝𝑀). Since the Hamiltonian function is invariant with respect to this canonical transformation, it follows, from (2.11) through (3.5), that 𝐻=𝐻0+𝐻𝐽2+𝐻𝛾,(3.8) where 𝐻0=𝑛𝑝𝑀,𝐻𝐽2=2𝑛𝑎𝜕𝑉2𝑝𝜕𝑀𝑎+1𝑒2𝑛𝑎2𝑒𝜕𝑉2+𝜕𝜔1𝑒2𝜕𝑉2𝑝𝜕𝑀𝑒+1𝑛𝑎21𝑒2sin𝐼𝜕𝑉2𝜕Ω+cos𝐼𝜕𝑉2𝑝𝜕𝜔𝐼+1𝑛𝑎21𝑒2sin𝐼𝜕𝑉2𝑝𝜕𝐼Ω+1𝑒2𝑛𝑎2𝑒𝜕𝑉2𝜕𝑒𝑒cot𝐼1𝑒2𝜕𝑉2𝑝𝜕𝐼𝜔+1𝑛𝑎2𝜕𝑉2𝜕𝑎1𝑒2𝑎𝑒𝜕𝑉2𝑝𝜕𝑒M,𝐻𝛾=12𝑛2𝑎21𝑒212(1cos2𝑓)2𝑎𝑒𝑝𝑎+1𝑒2𝑝𝑒2+21𝑒2sin2𝑓𝑎𝑝𝑎𝑝𝜔1𝑒2𝑝2𝑒𝑒𝑝𝜔+41𝑒23/2sin𝑓2𝑒1+𝑒cos𝑓+cos𝑓𝑎𝑝𝑎𝑝𝑀+1𝑒2𝑝2𝑒𝑒𝑝𝑀+1𝑒222𝑒2(1+cos2𝑓)𝑝𝜔221𝑒25/2𝑒22𝑒1+𝑒cos𝑓+cos𝑓cos𝑓𝑝𝜔𝑝𝑀+1𝑒23𝑒22𝑒1+𝑒cos𝑓+cos𝑓2𝑝𝑀2+4𝑎21𝑒22𝑎𝑟2𝑝𝑎2+4𝑎1𝑒22𝑎𝑟(cos𝐸+cos𝑓)𝑝𝑎𝑝𝑒+1𝑒22(cos𝐸+cos𝑓)2𝑝𝑒2+4𝑎1𝑒22𝑒𝑎𝑟1sin𝑓1+×𝑝1+𝑒cos𝑓𝑎𝑝𝜔1𝑒21/2𝑝𝑎𝑝𝑀+21𝑒22𝑒1(cos𝐸+cos𝑓)1+𝑝1+𝑒cos𝑓×sin𝑓𝑒𝑝𝜔1𝑒2𝑝𝑒𝑝𝑀+1𝑒2𝑒11+𝑝1+𝑒cos𝑓sin𝑓𝜔1𝑒2𝑝𝑀2+12𝑟𝑎2𝑝𝐼2+𝑝Ωsin𝐼𝑝𝜔cot𝐼2+12𝑟𝑎2𝑝cos2(𝜔+𝑓)𝐼2𝑝Ωsin𝐼𝑝𝜔cot𝐼2+𝑟𝑎2sin2(𝜔+𝑓)𝑝𝐼𝑝Ωsin𝐼𝑝𝜔,cot𝐼(3.9) with the disturbing function 𝑉2 given by 𝑉2=𝜇𝑎𝑎𝑒𝑎2𝐽2𝑎𝑟31234sin2𝐼+34𝑎𝑟3sin2𝐼cos2(𝜔+𝑓).(3.10)

The new Hamiltonian function 𝐻, defined by (3.8)–(3.10), describes the optimal low-thrust limited-power trajectories in a noncentral gravity field which includes the perturbative effects of the second zonal harmonic of the gravitational potential.

4. Averaged Maximum Hamiltonian for Optimal Transfers

In order to eliminate the short periodic terms from the maximum Hamiltonian function 𝐻, Hori method [29] is applied. It is assumed that 𝐻0 is of zero-order and 𝐻𝛾 is of the first-order in a small parameter associated to the magnitude of the thrust acceleration, and 𝐻𝐽2 has the same order of 𝐻𝛾.

Consider an infinitesimal canonical transformation: 𝑎,𝑒,𝐼,Ω,𝜔,𝑀,𝑝𝑎,𝑝𝑒,𝑝𝐼,𝑝Ω,𝑝𝜔,𝑝𝑀𝑎,𝑒,𝐼,Ω,𝜔,𝑀,𝑝𝑎,𝑝𝑒,𝑝𝐼,𝑝Ω,𝑝𝜔,𝑝𝑀.(4.1) The new variables are designated by the prime. According to the algorithm of Hori method [29], at order 0, 𝐹0=𝑛𝑝𝑀.(4.2)𝐹0 denotes the new undisturbed Hamiltonian. Now, consider the canonical system described by 𝐹0: 𝑑𝑎𝑑𝑡=0,𝑑𝑒𝑑𝑡=0,𝑑𝐼𝑑𝑡=0,𝑑Ω𝑑𝑡=0,𝑑𝜔𝑑𝑡=0,𝑑𝑀𝑑𝑡=𝑛,𝑑𝑝𝑎=3𝑑𝑡2𝑛𝑎𝑝𝑀,𝑑𝑝𝑒𝑑𝑡=0,𝑑𝑝𝐼𝑑𝑡=0,𝑑𝑝Ω𝑑𝑡=0,𝑑𝑝𝜔𝑑𝑡=0,𝑑𝑝𝑀𝑑𝑡=0,(4.3) general solution of which is given by 𝑎=𝑎0,𝑒=𝑒0,𝐼=𝐼0,Ω=Ω0,𝜔=𝜔0,𝑀=𝑀0+𝑛𝑡𝑡0,𝑝𝑎=𝑝𝑎0+32𝑛𝑡𝑡0𝑎𝑝𝑀,𝑝𝑒=𝑝𝑒0,𝑝𝐼=𝑝𝐼0,𝑝Ω=𝑝Ω0,𝑝𝜔=𝑝𝜔0,𝑝𝑀=𝑝𝑀0.(4.4) The subscript 0 denotes the constants of integration.

Introducing the general solution defined above into the equation of order 1 of the algorithm of Hori method, it reduces to 𝜕𝑆1𝜕𝑡=𝐻𝐽2+𝐻𝛾𝐹1.(4.5) According to [29], the mean value of 𝐻𝐽2+𝐻𝛾 must be calculated, and, 𝑆1 is obtained through integration of the remaining part. 𝐹1 and 𝑆1 are given by the following equations: 𝐹13=2𝑛𝐽2𝑎𝑒𝑎21𝑒22cos𝐼𝑝Ω+1215cos2𝐼𝑝𝜔+𝑎2𝜇4𝑎2𝑝𝑎2+521𝑒2𝑝𝑒2+54𝑒22𝑒2𝑝𝜔2+𝑝𝐼221𝑒231+2𝑒2+52𝑒2cos2𝜔+5𝑒2sin2𝜔21𝑒2𝑝𝐼𝑝Ωsin𝐼cot𝐼𝑝𝜔+121𝑒2𝑝Ωsin𝐼cot𝐼𝑝𝜔231+2𝑒252𝑒2cos2𝜔,𝑆1=𝐽2𝑎𝑒𝑎2312sin2𝐼𝑎𝑟31𝑒23/2+32sin2𝐼𝑎𝑟3𝜔cos2+𝑓𝑎𝑝𝑎+32sin2𝐼𝑒1𝑒212𝜔cos2+𝑓𝑒2cos2𝜔+𝑓+13cos2𝜔+3𝑓+1𝑒2𝑒1234sin2𝐼𝑎𝑟31𝑒23/2+34sin2𝐼𝑎𝑟3𝜔cos2+𝑓𝑝𝑒+34sin2𝐼1𝑒2212𝜔cos2+𝑓+𝑒2cos2𝜔+𝑓+13cos2𝜔+3𝑓𝑝𝐼+32cos𝐼1𝑒22𝑓𝑀+𝑒sin𝑓+12𝜔sin2+𝑓+𝑒2sin2𝜔+𝑓+13sin2𝜔+3𝑓𝑝Ω+345cos2𝐼11𝑒22𝑓𝑀+𝑒sin𝑓+143cos2𝐼1𝑒1𝑒2𝑎𝑟21𝑒2+𝑎𝑟+1sin𝑓+38sin2𝐼𝑒1𝑒2𝑎𝑟21𝑒2𝑎𝑟+1sin2𝜔+𝑓+𝑎𝑟21𝑒2+𝑎𝑟+13sin2𝜔+3𝑓+3835cos2𝐼1𝑒22𝜔sin2+𝑓+𝑒sin2𝜔+𝑓+13sin2𝜔+3𝑓𝑝𝜔+12𝑎5𝜇38𝑒sin𝐸𝑎2𝑝𝑎2+81𝑒2sin𝐸𝑎𝑝𝑎𝑝𝑒81𝑒2𝑒cos𝐸𝑝𝑎𝑝𝜔+1𝑒254𝑒sin𝐸+34sin2𝐸1𝑒12sin3𝐸𝑝𝑒2+1𝑒2𝑒52𝑒cos𝐸123𝑒2cos2𝐸+16𝑒cos3𝐸𝑝𝑒𝑝𝜔+1𝑒21𝑝𝐼2+𝑝Ωsin𝐼𝑝𝜔cot𝐼2×𝑒+38𝑒3sin𝐸+38𝑒2sin2𝐸1𝑒243sin3𝐸+1𝑒21×𝑝𝐼2cos2𝜔+2𝑝𝐼𝑝Ωsin𝐼𝑝𝜔cot𝐼sin2𝜔𝑝Ωsin𝐼𝑝𝜔cot𝐼2cos2𝜔×54𝑒+58𝑒3sin𝐸+14+18𝑒2sin2𝐸+1𝑒12+1𝑒243sin3𝐸+1𝑒21/2𝑝𝐼2sin2𝜔+2𝑝𝐼𝑝Ωsin𝐼𝑝𝜔cot𝐼cos2𝜔+𝑝Ωsin𝐼𝑝𝜔cot𝐼2sin2𝜔×54𝑒cos𝐸14+14𝑒2cos2𝐸+1𝑒12cos3𝐸+𝑝𝜔2𝑒254𝑒𝑒3sin𝐸+34+12𝑒2sin2𝐸+1𝑒12sin3𝐸.(4.6) Terms factored by 𝑝𝑀 have been omitted in equations above, since only transfers (no rendez-vous) are considered [2].

It should be noted that the averaged maximum Hamiltonian and the generating function become singular for circular and/or equatorial orbits. In order to avoid these singularities, suitable sets of nonsingular elements will be introduced in the next sections.

5. Optimal Transfers between Close Quasicircular Orbits

In this section, approximate first-order analytical solutions will be obtained for transfers between close quasicircular orbits.

5.1. Transfers between Close Quasicircular Nonequatorial Orbits

Consider the Mathieu transformation [33] defined by 𝑎=𝑎,=𝑒cos𝜔,𝑘=𝑒sin𝜔,𝐼=𝐼,Ω=Ω,=𝑀+𝜔,𝑝(5.1)𝑎=𝑝𝑎,𝑝=𝑝𝑒cos𝜔𝑝𝜔𝑝𝑀𝑒sin𝜔,𝑝𝑘=𝑝𝑒sin𝜔+𝑝𝜔𝑝𝑀𝑒cos𝜔,𝑝𝐼=𝑝𝐼,𝑝Ω=𝑝Ω,𝑝=𝑝𝑀.(5.2) The new set of canonical variables, designated by the double prime, is nonsingular for circular nonequatorial orbits.

Introducing (5.2) into (4.6) and using the expansions of the elliptic motion in terms of eccentricity and mean anomaly [31], one finds that the new averaged maximum Hamiltonian and the generating function are written up to the zeroth order in eccentricity as 𝐹13=2𝑛𝐽2𝑎𝑒𝑎2cos𝐼𝑝Ω+𝑎2𝜇4𝑎2𝑝2𝑎+52𝑝2+𝑝2𝑘+12𝑝2𝐼+𝑝2Ωsin2𝐼,𝑆1=32𝐽2𝑎𝑒𝑎2𝑎𝑝𝑎sin2𝐼cos2+𝑝1cos+4sin2𝐼75cos+3cos3+𝑝𝑘1sin+4sin2𝐼77sin+3+1sin34𝑝𝐼1sin2𝐼cos2+2𝑝Ω+1cos𝐼sin22𝑎5𝜇38𝑎𝑝𝑎𝑝sin𝑝𝑘1cos23𝑝𝑝𝑘+𝑝Ω𝑝sin𝐼𝐼+1cos243𝑝2𝑝2𝑘+𝑝2𝐼𝑝2Ωsin2𝐼.sin2(5.3) Double prime is omitted to simplify the notation.

For transfers between close orbits, 𝐹1 and 𝑆1 can be linearized around a suitable reference orbit, and an approximate first-order analytical solution can be determined. This solution is given by Δ𝑥=𝐴𝑝0+𝐵,(5.4) where Δ𝑥=[Δ𝛼ΔΔ𝑘Δ𝐼ΔΩ]𝑇,𝑝0 is the 5×1 vector of initial value of the adjoint variables, 𝐴 is a 5×5 symmetric matrix concerning the optimal thrust acceleration and 𝐵 is a 5×1 vector containing the perturbative effects of the oblateness of the central body. The variables 𝛼=𝑎/𝑎 and 𝑝𝛼=𝑎𝑝𝑎 are introduced to make the adjoint and the state vectors dimensionless, and the variables Ω=Ωsin𝐼 and 𝑝Ω=𝑝Ω/sin𝐼 are introduced to simplify the matrix 𝐴. The adjoint vector is constant. The matrix 𝐴 and the vector 𝐵 are given by 𝑎𝐴=𝛼𝛼𝑎𝛼𝑎𝛼𝑘𝑎00𝛼𝑎𝑎𝑘𝑎00𝑘𝛼𝑎𝑘𝑎𝑘𝑘00000𝑎𝐼𝐼𝑎𝐼Ω000𝑎Ω𝐼𝑎ΩΩ𝑏,(5.5)𝐵=𝛼𝑏𝑏𝑘𝑏𝐼𝑏Ω,(5.6) where 𝑎𝛼𝛼=4𝑎5𝜇3|||||𝑓0,𝑎𝛼=𝑎𝛼=4𝑎5𝜇3sin|||||𝑓0,𝑎𝛼𝑘=𝑎𝑘𝛼=4𝑎5𝜇3cos|||||𝑓0,𝑎=𝑎5𝜇3523+4sin2|||||𝑓0,𝑎𝑘=𝑎𝑘3=4𝑎5𝜇3cos2|||||𝑓0,𝑎𝑘𝑘=𝑎5𝜇35234sin2|||||𝑓0,𝑎𝐼𝐼=𝑎5𝜇3121+4sin2|||||𝑓0,𝑎𝐼Ω=𝑎Ω𝐼1=4𝑎5𝜇3cos2|||||𝑓0,𝑎ΩΩ=𝑎5𝜇31214sin2|||||𝑓0,𝑏𝛼=𝜀sin2𝐼cos2|||𝑓0,𝑏=𝜀cos1+4sin2𝐼5cos7+3cos3|||𝑓0,𝑏𝑘=𝜀sin1+4sin2𝐼7sin7+3sin3|||𝑓0,𝑏𝐼=𝜀4sin2𝐼cos2|||𝑓0,𝑏Ω𝜀=2sin2𝐼12sin2|||𝑓0,(5.7) with 𝑓=0+𝑛(𝑡𝑓𝑡0),𝑡0 is the initial time, 𝑡𝑓 is the final time, and 𝜀=(3/2)𝐽2(𝑎𝑒/𝑎)2. The overbar denotes the orbital elements of the reference orbit.

The analytical solution defined by (5.4)–(5.7) is in agreement with the ones obtained through different approaches in [6, 10]. In [6] the optimization problem is formulated with Gauss equations in nonsingular orbital elements as state equations, and in [10] Hori method is applied after the transformation of variables. Equations (5.4)–(5.7) represent a complete first-order analytical solution for optimal low-thrust limited-power transfers between close quasicircular nonequatorial orbits in a gravity field that includes the second zonal harmonic 𝐽2 in the gravitational potential. These equations contain arbitrary constants of integration that must be determined to satisfy the two-point boundary value problem of going from the initial orbit at the time 𝑡0=0 to the final orbit at the specified final time 𝑡𝑓=𝑇. Since they are linear in these constants, the boundary value problem can be solved by simple techniques.

An approximate expression for the optimal thrust acceleration 𝜸 can be obtained from (3.5) and (5.2), by using the expansions of the elliptic motion, and it is given, up to the zeroth order in eccentricity, by 𝜸=1𝑝𝑛𝑎sin𝑝𝑘𝐞cos𝑟𝑝+2𝛼+𝑝cos+𝑝𝑘𝐞sin𝑠+𝑝𝐼cos+𝑝Ω𝐞sin𝑤.(5.8)

According to Section 2, the optimal fuel consumption variable 𝐽 is determined by simple quadrature of the last differential equation in (2.3), with 𝜸 given by (5.8). An approximate expression for 𝐽 is given by 1Δ𝐽=2𝑎𝛼𝛼𝑝2𝛼+2𝑎𝛼𝑝𝛼𝑝+2𝑎𝛼𝑘𝑝𝛼𝑝𝑘+𝑎𝑝2+2𝑎𝑘𝑝𝑝𝑘+𝑎𝑘𝑘𝑝2𝑘+𝑎𝐼𝐼𝑝2𝐼+2𝑎Ω𝑝𝐼𝑝Ω+𝑎ΩΩ𝑝2Ω.(5.9)

For transfers involving a large number of revolutions, (5.4)–(5.7) can be greatly simplified by neglecting the short periodic terms in comparison to the secular ones, and, the system of algebraic can be solved analytically. Thus, 𝑝𝛼=14𝜇3𝑎5Δ𝛼Δ,𝑝=25𝜇3𝑎5ΔΔ,𝑝𝑘=25𝜇3𝑎5Δ𝑘Δ,𝑝𝐼=2𝜇3𝑎5Δ𝐼Δ,𝑝Ω=2𝜇3𝑎5ΔΩsin𝐼Δ+𝜀2sin2𝐼.(5.10) Introducing these equations into (5.8) and (5.9), 𝜸 and 𝐽 are obtained explicitly in terms of the imposed variations of the orbital elements, the perturbative effects due to the oblateness of the central body and the duration of the maneuver: 𝜸=𝜇𝑎225ΔΔsinΔ𝑘Δcos𝐞𝑟+12Δ𝛼Δ+45ΔΔcos+Δ𝑘Δsin𝐞𝑠+2Δ𝐼Δcos+ΔΩΔ+𝜀cos𝐼sin𝐼sin𝐞𝑤,(5.11)Δ𝐽Δ=14𝜇3𝑎512Δ𝛼Δ2+45Δ2+Δ𝑘2Δ2+4Δ𝐼Δ2+ΔΩΔ+𝜀cos𝐼2sin2𝐼.(5.12)

5.2. Transfers between Close Quasicircular Near-Equatorial Orbits

Consider the Mathieu transformation [33] defined by 𝑎=𝑎,𝜉=𝑒𝜔cos+Ω,𝜂=𝑒𝜔sin+Ω,𝑃𝐼=sin2cosΩ,𝑄𝐼=sin2sinΩ,𝜆=𝑀+𝜔+Ω,𝑝(5.13)𝑎=𝑝𝑎,𝑝𝜉=𝑝𝑒𝜔cos+Ω𝑝𝜔𝑝𝑀𝑒𝜔sin+Ω,𝑝𝜂=𝑝𝑒𝜔sin+Ω+𝑝𝜔𝑝𝑀𝑒𝜔cos+Ω,𝑝𝑃=𝑝𝐼2cosΩ𝐼cos𝑝/2Ω𝑝𝜔sinΩ𝐼sin,𝑝/2𝑄=𝑝𝐼2sinΩ𝐼cos+𝑝/2Ω𝑝𝜔cosΩ𝐼sin,𝑝/2𝜆=𝑝𝑀.(5.14) The new set of canonical variables, designated by the double prime, is nonsingular for circular equatorial orbits.

Introducing (5.14) into (4.6) and using the expansions of the elliptic motion in terms of eccentricity and mean anomaly [31], one finds that the new averaged maximum Hamiltonian and the generating function are written up to the zeroth order in eccentricity and first-order in inclination as 𝐹1=32𝑛𝐽2𝑎𝑒𝑎2𝑄𝑝𝑃𝑃𝑝𝑄+𝑎2𝜇4𝑎2𝑝2𝑎+52𝑝2𝜉+𝑝2𝜂+18𝑝2𝑃+𝑝2𝑄,𝑆(5.15)1=32𝐽2𝑎𝑒𝑎2𝑝𝜉cos𝜆+𝑝𝜂1sin𝜆+2𝑄𝑝𝑃+𝑃𝑝𝑄1sin2𝜆+2𝑃𝑝𝑃𝑄𝑝𝑄+1cos2𝜆2𝑎5𝜇38𝑎𝑝𝑎𝑝𝜉sin𝜆𝑝𝜂+3cos𝜆4𝑝2𝜉𝑝2𝜂+1𝑝162𝑃𝑝2𝑄3sin2𝜆2𝑝𝜉𝑝𝜂+18𝑝𝑃𝑝𝑄.cos2𝜆(5.16) Double prime is omitted to simplify the notation.

For transfers between close orbits, 𝐹1 and 𝑆1 can be linearized around a suitable reference orbit, and an approximate first-order analytical solution can be determined. This solution is given by Δ𝑥=𝐶𝑝0+𝐷,(5.17) where Δ𝑥=[Δ𝛼Δ𝜉Δ𝜂Δ𝑃Δ𝑄]𝑇,𝑝0 is the 5×1 vector of initial value of the adjoint variables, 𝐶 is a 5×5 matrix concerning the optimal thrust acceleration, and 𝐷 is a 5×1 vector containing the perturbative effects of the oblateness of the central body. The matrix 𝐶 and the vector 𝐷 are given by 𝑐𝐶=𝛼𝛼𝑐𝛼𝜉𝑐𝛼𝜂𝑐00𝜉𝛼𝑐𝜉𝜉𝑐𝜉𝜂𝑐00𝜂𝛼𝑐𝜂𝜉𝑐𝜂𝜂00000𝑐𝑃𝑃𝑐𝑃𝑄000𝑐𝑄𝑃𝑐𝑄𝑄0𝑑,(5.18)𝐷=𝜉𝑑𝜂𝑑𝑃𝑑𝑄,(5.19) where 𝑐𝛼𝛼=4𝑎5𝜇3𝜆|||||𝜆𝑓𝜆0,𝑐𝛼𝜉=𝑐𝜉𝛼=4𝑎5𝜇3sin𝜆|||||𝜆𝑓𝜆0,𝑐𝛼𝜂=𝑐𝜂𝛼=4𝑎5𝜇3cos𝜆|||||𝜆𝑓𝜆0,𝑐𝜉𝜉=𝑎5𝜇3523𝜆+4sin2𝜆|||||𝜆𝑓𝜆0,𝑐𝜉𝜂=𝑐𝜂𝜉3=4𝑎5𝜇3cos2𝜆|||||𝜆𝑓𝜆0,𝑐𝜂𝜂=𝑎5𝜇3523𝜆4sin2𝜆|||||𝜆𝑓𝜆0,𝑐𝑃𝑃=18𝑎5𝜇3Δ𝜆cos𝜀Δ1𝜆+2sin(2+𝜀)𝜆𝑓𝜀𝜆012sin(2+𝜀)𝜆0𝜀𝜆𝑓,𝑐𝑃𝑄=18𝑎5𝜇3Δ𝜆sin𝜀Δ1𝜆2cos(2+𝜀)𝜆𝑓𝜀𝜆0+12cos(2+𝜀)𝜆0𝜀𝜆𝑓,𝑐𝑄𝑃=18𝑎5𝜇3Δ𝜆sin𝜀Δ1𝜆2cos(2+𝜀)𝜆𝑓𝜀𝜆0+12cos(2+𝜀)𝜆0𝜀𝜆𝑓,𝑐𝑄𝑄=18𝑎5𝜇3Δ𝜆cos𝜀Δ1𝜆2(sin2+𝜀)𝜆𝑓𝜀𝜆0+12(sin2+𝜀)𝜆0𝜀𝜆𝑓,𝑑𝜉=32𝐽2𝑎𝑒𝑎2cos𝜆||||𝜆𝑓𝜆0,𝑑𝜂=32𝐽2𝑎𝑒𝑎2sin𝜆||||𝜆𝑓𝜆0,𝑑𝑃=𝑃0cos𝜀Δ1𝜆2𝜀𝜀cos𝜆𝑓(2+𝜀)𝜆0+12𝜀𝜀cos𝜆0(2+𝜀)𝜆𝑓1+𝑄0sin𝜀Δ1𝜆+2𝜀𝜀sin𝜆𝑓(2+𝜀)𝜆012𝜀𝜀sin𝜆0(2+𝜀)𝜆𝑓,𝑑𝑄=𝑄0cos𝜀Δ1𝜆+2𝜀𝜀cos𝜆𝑓(2+𝜀)𝜆012𝜀𝜀cos𝜆0(2+𝜀)𝜆𝑓1+𝑃0sin𝜀Δ1𝜆+2𝜀𝜀sin𝜆𝑓(2+𝜀)𝜆012𝜀𝜀sin𝜆0(2+𝜀)𝜆𝑓.(5.20) The adjoint variables are 𝑝𝑎=𝑝𝑎0,𝑝𝜉=𝑝𝜉0,𝑝𝜂=𝑝𝜂0,𝑝𝑃=𝑝𝑃0cos𝜀Δ𝜆+𝑝𝑄0sin𝜀Δ𝑝𝜆,𝑄=𝑝𝑃0sin𝜀Δ𝜆+𝑝𝑄0cos𝜀Δ𝜆,(5.21) where Δ𝜆=𝜆𝑓𝜆0=𝑛(𝑡𝑓𝑡0). The overbar denotes the orbital elements of the reference orbit. Equations (5.20) do not contain higher order terms in 𝜀  as appear implicitly in the solution presented in [7].

The analytical solution defined by (5.17)–(5.21) is in agreement with the ones obtained through different approaches in [7, 10]. In [7] the optimization problem is formulated with Gauss equations in nonsingular orbital elements as state equations, and in [10] Hori method is applied after the transformation of variables. Equations (5.17)–(5.21) represent a complete first-order analytical solution for optimal low-thrust limited-power transfers between close quasicircular near-equatorial orbits in a gravity field that includes the second zonal harmonic 𝐽2 in the gravitational potential. These equations contain arbitrary constants of integration that must be determined to satisfy the two-point boundary value problem of going from the initial orbit at the time 𝑡0=0 to the final orbit at the specified final time 𝑡𝑓=𝑇. Since they are linear in these constants, the boundary value problem can be solved by simple techniques.

As described in Section 5.1, approximate expressions can be obtained for the optimal thrust acceleration 𝜸 and the fuel consumption 𝐽. These expressions are given by 𝜸=1𝑝𝑛𝑎𝜉sin𝜆𝑝𝜂𝐞cos𝜆𝑟𝑝+2𝛼+𝑝𝜉cos𝜆+𝑝𝜂𝐞sin𝜆𝑠+12𝑝𝑃cos𝜆+𝑝𝑄𝐞sin𝜆𝑤,1Δ𝐽=2𝑐𝛼𝛼𝑝2𝛼+2𝑐𝛼𝜉𝑝𝛼𝑝𝜉+2𝑐𝛼𝜂𝑝𝛼𝑝𝜂+𝑐𝜉𝜉𝑝2𝜉+2𝑐𝜉𝜂𝑝𝜉𝑝𝜂+𝑐𝜂𝜂𝑝2𝜂+𝑐𝑃𝑃𝑝2𝑃+𝑐𝑃𝑄+𝑐𝑄𝑃𝑝𝑃𝑝𝑄+𝑐𝑄𝑄𝑝2𝑄.(5.22)

For transfers involving a large number of revolutions, (5.17)–(5.21) can be greatly simplified by neglecting the short periodic terms in comparison to the secular ones, and, the system of algebraic can be solved analytically. Thus, 𝑝𝛼=14𝜇3𝑎5Δ𝛼Δ𝜆,𝑝𝜉=25𝜇3𝑎5Δ𝜉Δ𝜆,𝑝𝜂=25𝜇3𝑎5Δ𝜂Δ𝜆,𝑝𝑃=𝜇3𝑎58Δ𝜆𝑃0+Δ𝑃cos𝜀Δ𝜆𝑄0+Δ𝑄sin𝜀Δ𝜆𝑃0,𝑝𝑄=𝜇3𝑎58Δ𝜆𝑄0+Δ𝑄cos𝜀Δ𝜆+𝑃0+Δ𝑃sin𝜀Δ𝜆𝑄0.(5.23) The subscript denoting the constants is omitted. Introducing these equations into (5.22), 𝜸 and 𝐽 are obtained explicitly in terms of the imposed variations of the orbital elements, the perturbative effects due to the oblateness of the central body and the duration of the maneuver: 𝜸=𝜇𝑎225Δ𝜉Δ𝜆sin𝜆Δ𝜂Δ𝜆cos𝜆𝐞𝑟+12Δ𝛼Δ𝜆+45Δ𝜉Δ𝜆cos𝜆+Δ𝜂Δ𝜆sin𝜆𝐞𝑠+4Δ𝜆𝑃𝑓cos𝜆+𝜃𝑓𝑃0cos𝜆+𝜃0+𝑄𝑓sin𝜆+𝜃𝑓𝑄0sin𝜆+𝜃0𝐞𝑤,Δ𝐽Δ𝜆=14𝜇3𝑎512Δ𝛼Δ𝜆2+45Δ𝜉2+Δ𝜂2Δ𝜆2+16cos𝜀Δ𝜆Δ𝑃2+Δ𝑄2Δ𝜆2+4Δ𝜆2sin𝜀Δ𝜆2×𝑃𝑓𝑃0sin𝜀Δ𝜆2𝑄0cos𝜀Δ𝜆2+𝑄𝑓𝑃0cos𝜀Δ𝜆2+𝑄0sin𝜀Δ𝜆2,(5.24) with 𝜃0=𝜀𝑛(𝑡𝑡0) and 𝜃𝑓=𝜀𝑛(𝑡𝑡𝑓).

5.3. General Remarks on Optimal Transfers between Close Quasicircular Orbits

Considering the analytical solutions described above, general remarks about the effects of the oblateness of the central body on optimal transfers between close quasicircular orbits can be stated.

(1) Matrices 𝐴 and 𝐶, defined by (5.5) and (5.18), respectively, can be decomposed into two square matrices 3×3 and 2×2. This fact shows that there exists an uncoupling between the in-plane modifications and the rotation of the orbital plane. Similar results were obtained by Edelbaum [1] and Marec [3] for transfers in a Newtonian central field.(2) There is a normal component of the optimal thrust acceleration that counteracts the perturbative effects due to the oblateness of the central body. This normal component is proportional to 𝜀.(3) In general, the maneuvers in a noncentral gravity field are more expensive than the maneuvers in Newtonian central field, taking into account that the perturbative effects due to the oblateness of the central body must be counteracted.(4) Fuel can be saved for long-time maneuvers which involves modification of the longitude of the ascending node if the terminal orbits are direct (0<𝐼<90) and ΔΩ<0, and if the terminal orbits are retrograde (90<𝐼<180) and ΔΩ>0.(5) The extra consumption needed to counteract the perturbative effects due to the oblateness of the central body reaches its maximum for maneuvers between orbits with 45 or 135 of inclination and its minimum (null extra consumption) for maneuvers between equatorial or polar orbits (the gravity field is symmetric in these cases). (6) As the transfer time increases, the extra fuel consumption increases.(7) The extra fuel consumption is greater for transfers between low orbits.

6. Optimal Long-Time Transfers between Arbitrary Orbits

In this section the two-point boundary-value problem for long-time transfers between arbitrary orbits is formulated. In Section 7, this boundary-problem problem will be solved numerically through a shooting method, and the solution will be compared to the analytical ones in the case of orbits with small eccentricities.

Since the averaged maximum Hamiltonian, given by (4.6), becomes singular for circular and/or equatorial orbits, a set of nonsingular elements is introduced.

Consider the Mathieu transformation [33] defined by 𝑎=𝑎,𝜉=𝑒𝜔cos+Ω,𝜂=𝑒𝜔sin+Ω,𝑃𝐼=sin2cosΩ,𝑄𝐼=sin2sinΩ,𝑝(6.1)𝑎=𝑝𝑎,𝑝𝜉=𝑝𝑒𝜔cos+Ω𝑝𝜔𝑝𝑀𝑒𝜔sin+Ω,𝑝𝜂=𝑝𝑒𝜔sin+Ω+𝑝𝜔𝑝𝑀𝑒𝜔cos+Ω,𝑝𝑃=𝑝𝐼2cosΩ𝐼cos𝑝/2Ω𝑝𝜔sinΩ𝐼sin,𝑝/2𝑄=𝑝𝐼2sinΩ𝐼cos+𝑝/2Ω𝑝𝜔cosΩ𝐼sin./2(6.2) This transformation of variables is the same one given by (5.12), excepting the fast-phase, which is unnecessary (short-periodic terms have been eliminated). Double prime designates the new variables.

Introducing (6.2) into (4.6), 𝐹1 is written as 𝐹13=2𝑛𝐽2𝑎𝑒𝑎21𝜉2𝜂22𝑃122+𝑄2𝜉𝑝𝜂𝜂𝑝𝜉+𝑃𝑝𝑄𝑄𝑝𝑃+12𝑃15122+𝑄22𝜉𝑝𝜂𝜂𝑝𝜉+𝑎2𝜇4𝑎2𝑝2𝑎+𝑝2𝜉5252𝜉22𝜂2+𝑝2𝜂5252𝜂22𝜉2𝜉𝜂𝑝𝜉𝑝𝜂+121𝜉2𝜂2131+2𝜉2+32𝜂2×14𝑝2𝑃+𝑝2𝑄14𝑃𝑝𝑃+𝑄𝑝𝑄2+141𝑃2𝑄21×𝑄𝑝𝑃𝑃𝑝𝑄2+4𝜉𝑝𝜂𝜂𝑝𝜉𝑃𝑝𝑄𝑄𝑝𝑃𝑃+42+𝑄2𝜉𝑝𝜂𝜂𝑝𝜉2581𝑃2𝑄21𝜉2𝜂2𝑃2𝑄2×2+4𝜉𝜂𝑃𝑄𝜉𝑝𝜂𝜂𝑝𝜉+𝑃𝑝𝑄𝑄𝑝𝑃2+1𝑃2𝑄2𝑝2𝑃+𝑝2𝑄52𝜉𝜂𝑝𝑃𝑝𝑄+58𝜉2𝜂2𝑝2𝑃𝑝2𝑄+52𝜉𝑝𝜂𝜂𝑝𝜉𝜉2𝜂2𝑃𝑝𝑄+𝑄𝑝𝑃+2𝜉𝜂𝑃𝑝𝑃𝑄𝑝𝑄.(6.3) Double prime is omitted to simplify the notation. Note that (5.15) is a simplification of (6.3) for quasicircular near-equatorial orbits.

The two-point boundary-value problem for long-time transfers between arbitrary orbits is formulated by the following system of canonical differential equations: 𝑑𝑥=𝑑𝑡𝜕𝐹1𝜕𝑝𝑇,𝑑𝑝𝑑𝑡=𝜕𝐹1𝜕𝑥𝑇,(6.4) where 𝑥=[𝑎𝜉𝜂𝑃𝑄]𝑇 and 𝑝=[𝑝𝑎𝑝𝜉𝑝𝜂𝑝𝑃𝑝𝑄]𝑇, and the boundary-conditions 𝑎𝑡0=𝑎0𝑡,𝜉0=𝜉0𝑡,𝜂0=𝜂0𝑡,𝑃0=𝑃0𝑡,𝑄0=𝑄0,𝑎𝑡𝑓=𝑎𝑓𝑡,𝜉𝑓=𝜉𝑓𝑡,𝜂𝑓=𝜂𝑓𝑡,𝑃𝑓=𝑃𝑓𝑡,𝑄𝑓=𝑄𝑓,(6.5) with initial time 𝑡0=0 and final time 𝑡𝑓=𝑇. The explicit form of state and adjoint equations is derived by using MAPLE Software 9, and they are not presented in text because they are extensive.

7. Results

In this section, numerical and analytical results are compared for maneuvers described in Tables 1 and 2. Earth is the central body, such that 𝐽2=1.0826×103. In Table 1, orbital elements of initial and final orbits are defined for a maneuver of modification of all orbital elements considering different values of transfer durations 𝑇=𝑡𝑓𝑡0 and ratios 𝑎𝑒/𝑎0. In Table 2, orbital elements of initial orbit, values of transfer duration and ratio 𝜌=𝑦𝑓/𝑦0, where 𝑦 denotes the orbital element that is changed in the maneuver, are defined for small (orbit corrections) and moderate amplitude maneuvers of modification of semimajor axis and inclination of the orbital plane. Results are presented in Figures 3 through 8 using canonical units. In these units, 𝑎0=1.0 and 𝜇=1.0. For example, in metric units, taking 𝑎𝑒=6378.2km,𝜌=0.950 corresponds to 𝑎0=6713.89km, and 𝑇=100 (time units) corresponds to 24.20 hours (1.008 day), and 𝜌=0.150 corresponds to 𝑎0=42521.34km, and 𝑇=250 (time units) corresponds to 964.45 hours (40.185 days). Thus, maneuvers described in the tables below have duration of 1 to 74 days, approximately. On the other hand, considering that the value of |𝜸| is approximately given by 2Δ𝐽/𝑇 (short periodic terms are neglected), |𝜸| has values between 1.0×104 and 4.0×103 (canonical units) for all results presented in Figures 35. These values of magnitude of the thrust acceleration are typical for LP systems [5, 27].

Figure 3 shows the consumption 𝐽 as function of ratio 𝑎𝑒/𝑎0 and values of 𝑇 for the maneuver of modification of all orbital elements described in Table 1. Figures 4 and 5 show the consumption 𝐽 as function of ratio 𝜌=𝑦𝑓/𝑦0 and 𝑇 for the small and moderate amplitude maneuvers described in Table 2, considering that the maneuvers are performed in central or noncentral gravity field. In these figures, solid line represents analytical solution, and dashed line represents numerical solution. Analytical solutions, including short periodic terms, are obtained by solving the linear system of algebraic equations given by (5.4)–(5.7) (transfer between nonequatorial orbits) with the semimajor axis and the inclination of the reference orbit defined by 𝑎=(𝑎0+𝑎𝑓)/2 and 𝐼=(𝐼0+𝐼𝑓)/2, respectively. In order to include the short periodic terms, it is assumed that the initial position of the vehicle is the pericenter of the initial orbit. Variables 𝐼 and Ω are transformed into nonsingular variables 𝑃 and 𝑄 to compare the analytical results with the numerical results. Numerical solutions are obtained by solving the two-point boundary value problem described by (6.4) and (6.5) through a shooting method [34].

Considering the maneuvers described in Tables 1 and 2, Figures 35 show that analytical and numerical solutions yield quite similar values of the consumption 𝐽 in the following cases: (i) maneuvers of modification of all orbital elements with ratio 𝑎𝑒/𝑎0<0.550; (ii) maneuvers of modification of semimajor axis with ratios 𝜌<1.25 and 𝑎𝑒/𝑎0<0.550; (iii) maneuvers of modification of inclination with ratios 𝜌<1.50 and 𝑎𝑒/𝑎0<0.550, for all values of 𝑇. For maneuvers with very high ratio 𝑎𝑒/𝑎0, analytical and numerical solutions only yield similar results for the same maneuvers described above for 𝑇=100 (time units).

Figures 68 represent the time history of state variables (nonsingular orbital elements) for maneuvers of modification of all orbital elements (Table 1) considering three values of ratios 𝑎𝑒/𝑎0—0.150, 0.550, and 0.950—and two values of transfer duration 𝑇—100 and 250 (time units). According to these figures, the analytical solution yields a good representation of the numerical solution for all state variables, excepting the semimajor axis for the maneuver with 𝑎𝑒/𝑎0=0.950 and 𝑇=250. The numerical solution shows that the semimajor axis is strongly perturbed by the oblateness of earth, and its time evolution cannot be described by the linear approximation given by the analytical solution. On the other hand, the analytical solutions shows that the amplitude of the short periodic terms decreases as the transfer duration increases such that these terms can be omitted for very long-time transfers.

8. Conclusion

In this paper an approach based on canonical transformations is presented for the problem of optimal low-thrust limited-power maneuvers between close orbits with small eccentricities in noncentral gravity which includes the perturbative effects of the second zonal harmonic of the gravitational potential. Analytical first-order solutions are derived through this approach for transfers between nonequatorial orbits and for transfers between near-equatorial orbits. These analytical solutions have been compared to a numerical solution for long-time transfers between arbitrary orbits described by averaged maximum Hamiltonian, expressed in nonsingular orbital elements. The results show that the analytical solutions yield good estimates of the fuel consumption for preliminary mission analysis considering maneuvers with moderate amplitude, ratio 𝑎𝑒/𝑎0 and transfers duration. Numerical results also show that the semimajor axis is strongly perturbed by the oblateness of earth, and its time evolution cannot be described by the linear approximation given by the analytical solution.

Appendix

The general solution of the differential equations for the adjoint variables 𝐩𝑟 and 𝐩𝑣 is obtained by computing the inverse of the Jacobian matrix of the point transformation between the Cartesian elements and the orbital ones, defined by (3.2) and (3.3). This matrix is obtained through the variations of the orbital elements induced by the variations in the Cartesian elements, as described below.

Let us consider the inverse of the point transformation defined by (3.2) and (3.3): 𝑟𝑎=2𝑟𝑣2,𝑒/𝜇2=12,𝜇𝑎cos𝐼=𝐤𝐡,cosΩ=𝐢𝐍𝑁,cos𝜔=𝐍𝐞,1𝑁𝑒cos𝐸=𝑒𝑟1𝑎,(A.1) where the eccentricity vector 𝐞, the angular momentum vector 𝐡, and the nodal vector 𝐍, shown in Figure 9, are given, as function of the Cartesian elements 𝐫 and 𝐯, by the following equations: 1𝐞=𝜇𝑣2𝜇𝑟,𝐫(𝐫𝐯)𝐯(A.2)𝐡=𝐫×𝐯,(A.3)𝐍=𝐤×𝐡.(A.4) Here the symbol × denotes the cross product. Note that the true anomaly 𝑓 has been replaced by eccentric anomaly 𝐸. These anomalies are related through the following equation: 𝑓tan2=1+𝑒𝐸1𝑒tan2.(A.5)

Now, consider the variations in the Cartesian elements, 𝛿𝐫 and 𝛿𝐯, given in the moving frame of reference by 𝛿𝐫=𝛿𝜉𝐞𝐫+𝛿𝜂𝐞𝑠+𝛿𝜁𝐞𝑤,𝛿𝐯=𝛿𝑢𝐞𝐫+𝛿𝑣𝐞𝑠+𝛿𝑤𝐞𝑤.(A.6) The variations of the orbital elements—𝑎,𝑒,𝐼,Ω,𝜔, and 𝐸—induced by the variations in the Cartesian elements, 𝛿𝐫 and 𝛿𝐯, are obtained straightforwardly from (A.1) through (A.6) and are given by 𝑎𝛿𝑎=2𝑟2𝐫𝛿𝐫𝑟+2𝑎2𝜇𝐯𝛿𝐯,𝛿𝑒=𝐡𝛿𝐡+𝜇𝑎𝑒22𝜇𝑎2𝑒𝛿𝑎,𝛿𝐼=𝐡𝛿𝐡2cot𝐼𝐤𝛿𝐡csc𝐼,𝛿Ω=𝐍𝛿𝐍𝑁2cotΩ𝐢𝛿𝐍𝑁cscΩ𝛿𝜔=𝛿𝑒𝑒+𝐍𝛿𝐍𝑁21cot𝜔[]1𝑁𝑒𝐍𝛿𝐞+𝐞𝛿𝐍csc𝜔,𝛿𝐸=𝑒sin𝐸𝐫𝛿𝐫𝑟𝑟𝑎𝑎2,𝛿𝑎+cos𝐸𝛿𝑒(A.7) where the variations of the vectors 𝐞,𝐡, and 𝐍 are written as 𝑣𝜇𝛿𝐞=2𝑠𝛿𝜉𝑣𝑠𝑣𝑟𝛿𝜂+2𝑟𝑣𝑠𝐞𝛿𝑣𝐫+𝑣𝑠𝑣𝑟𝑣𝛿𝜉+2𝑟𝜇𝑟𝛿𝜂𝑟𝑣𝑠𝛿𝑢𝑟𝑣𝑟𝐞𝛿𝑣𝑠+𝑣2𝜇𝑟𝛿𝜁𝑟𝑣𝑟𝐞𝛿𝑤𝑤,𝛿𝐡=𝑣𝑠𝛿𝜁𝐞𝑟+𝑣𝑟𝐞𝛿𝜁𝑟𝛿𝑤𝑠+𝑣𝑠𝛿𝜉𝑣𝑟𝐞𝛿𝜂+𝑟𝛿𝑣𝐰,𝑣𝛿𝐍=𝑠𝛿𝜉𝑣𝑟𝑣𝛿𝜂+𝑟𝛿𝑣sin𝐼cos(𝜔+𝑓)𝑟𝐞𝛿𝜁𝑟𝛿𝑤cos𝐼𝐫+𝑣𝑠𝛿𝜉𝑣𝑟𝛿𝜂+𝑟𝛿𝑣sin𝐼sin(𝜔+𝑓)𝑣𝑠𝐞cos𝐼𝛿𝜁𝑠+𝑣𝑟𝛿𝜁𝑟𝛿𝑤sin𝐼sin(𝜔+𝑓)+𝑣𝑠𝐞sin𝐼cos(𝜔+𝑓)𝛿𝜁𝑤.(A.8) Here 𝑣𝑟 and 𝑣𝑠 denote the radial and circumferential components of the velocity vector, respectively (see (3.3).

In the moving fame of reference, 𝐞,𝐡,𝐍, and the unit vectors 𝐢 and 𝐤 are written as 𝐞=𝑣𝑠𝜇𝐞1𝐫𝑣𝑟𝜇𝐞𝑠,(A.9)𝐡=𝐞𝐰=𝜇𝑎1𝑒2𝐞𝐰,(A.10)𝐍=sin𝐼cos(𝜔+𝑓)𝐞𝐫sin𝐼sin(𝜔+𝑓)𝐞𝑠,(A.11)𝐢=(cosΩcos(𝜔+𝑓)sinΩsin(𝜔+𝑓)cos𝐼)𝐞𝑟(cosΩsin(𝜔+𝑓)+sinΩcos(𝜔+𝑓)cos𝐼)𝐞𝑠+sinΩsin𝐼𝐞𝑤,(A.12)𝐤=sin(𝜔+𝑓)sin𝐼𝐞𝑟+cos(𝜔+𝑓)sin𝐼𝐞𝑠+cos𝐼𝐞𝑤.(A.13)

From (3.2), (3.3), and (A.2) through (A.11), one gets the explicit form of the variations of the orbital elements—𝑎,𝑒,𝐼,Ω,𝜔, and 𝐸—induced by the variations in the Cartesian elements, 𝛿𝐫 and 𝛿𝐯. At this point, it is useful to replace the variation in the eccentric anomaly 𝐸 by the variation in the mean anomaly 𝑀, obtained from the well-known Kepler’s equation 𝑀=𝐸𝑒sin𝐸.(A.14) This variation is given by 𝑟𝛿𝑀=𝑎𝛿𝐸sin𝐸𝛿𝑒.(A.15) Thus, 𝑎𝛿𝑎=2𝑟2𝛿𝜉+2𝑒sin𝑓𝑛1𝑒22𝛿𝑢+1𝑒2𝑛𝑎𝑟𝑎𝛿𝑣,𝛿𝑒=1𝑒2𝑟2cos𝐸𝛿𝜉+sin𝑓𝑎𝛿𝜂+1𝑒2𝑛𝑎sin𝑓𝛿𝑢+1𝑒21𝑛𝑎(cos𝐸+cos𝑓)𝛿𝑣,𝛿𝐼=𝑟cos𝐸sin𝜔+sin𝐸cos𝜔1𝑒21𝛿𝜁+𝑛𝑎1𝑒2𝑟𝑎1cos(𝜔+𝑓)𝛿𝑤,𝛿Ω=𝑟sin𝐼cos𝐸cos𝜔+sin𝐸sin𝜔1𝑒21𝛿𝜁+𝑛𝑎sin𝐼1𝑒2𝑟𝑎sin(𝜔+𝑓)𝛿𝑤,𝛿𝜔=sin𝑓𝑒𝑟𝛿𝜉𝑒+cos𝑓𝑎𝑒1𝑒2𝛿𝜂+cot𝐼𝑟cos𝐸cos𝜔sin𝐸sin𝜔1𝑒2𝛿𝜁1𝑒2𝑛𝑎𝑒cos𝑓𝛿𝑢+1𝑒21𝑛𝑎𝑒sin𝑓1+1+𝑒cos𝑓𝛿𝑣cot𝐼𝑛𝑎1𝑒2𝑟𝑎sin(𝜔+𝑓)𝛿𝑤,𝛿𝑀=1𝑒3cos𝐸𝑒𝑟1𝑒2sin𝑓𝛿𝜉+1𝑒2+𝑎𝑒cos𝑓𝛿𝜂1𝑒2𝑛𝑎𝑒cos𝑓2𝑒1+𝑒cos𝑓𝛿𝑢1𝑒21𝑛𝑎𝑒sin𝑓1+1+𝑒cos𝑓𝛿𝑣.(A.16)

Equations (A.16) can be put in the form 𝛿𝑎𝛿𝑒𝛿𝐼𝛿Ω𝛿𝜔𝛿𝑀=Δ1𝛿𝜉𝛿𝜂𝛿𝜁𝛿𝑢𝛿𝑣𝛿𝑤,(A.17) where the matrix Δ1 is inverse Jacobian matrix of the point transformation between the Cartesian elements and the orbital ones, defined by (3.2) and (3.3).

Following the properties of generalized canonical systems [32], the general solution of the differential equations for the adjoint variables 𝐩𝑟 and 𝐩𝑣 is given by 𝐩𝐫𝐩𝐯=Δ1𝑇𝑝𝑎𝑝𝑒𝑝𝐼𝑝Ω𝑝𝜔𝑝𝑀,(A.18) with 𝐩𝑟 and 𝐩𝑣 expressed in the moving frame of reference. Equations (3.4) and (3.5) are obtained straightforwardly from the above equation.

Acknowledgment

This research has been supported by CNPq under contract 305049/2006-2.