Next Article in Journal
The Andromeda Galaxy and Its Star Formation History
Previous Article in Journal
The ESSnuSB Design Study: Overview and Future Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Trajectory Optimization and Control Applied to Landing Maneuvers on Phobos from Mars-Phobos Distant Retrograde Orbits

Aerospace Engineering Department, College of Engineering, Embry-Riddle Aeronautical University, Daytona Beach, FL 32114, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Universe 2023, 9(8), 348; https://doi.org/10.3390/universe9080348
Submission received: 9 June 2023 / Revised: 18 July 2023 / Accepted: 21 July 2023 / Published: 25 July 2023
(This article belongs to the Section Gravitation)

Abstract

:
This paper presents research on the application of trajectory design, optimization, and control to an orbital transfer from Mars–Phobos Distant Retrograde Orbits to the surface of Phobos. Given a Distant Retrograde Orbit and a landing location on the surface of Phobos, landing trajectories for which total Δ v for a direct 2-burn maneuver is minimized are computed. This is accomplished through the use of Particle Swarm Optimization in which the required Δ v and time of flight are optimization parameters. The non-uniform gravitational environment of Phobos is considered in the computation. Results show how direct transfers can be achieved with Δ v on the order of ∼30 m/s.

1. Introduction

In order to facilitate the exploration of the Moon and Mars, intermediate staging locations have been proposed. The Lunar Gateway represents an intermediate orbital platform capable of linking, organizing, and supporting missions between the Earth and the Moon [1,2,3]. However, a lesser amount of literature regarding the use of similar platforms in the Martian environment exists. Such a Martian platform would be referred to as Mars Base Camp [4,5]. As well as the Lunar Gateway, Mars Base Camp is a proposed platform that would enable and aid the exploration of Mars and its surroundings, including its moons, Phobos and Deimos [6,7,8]. Although a final location for Mars Base Camp has not been chosen, it has been proposed that it could be located in the vicinity of Phobos or Deimos [9], such as a Mars–Phobos Distant Retrograde Orbit (MP DRO) [5]. This would avoid locating Mars Base Camp deep in Mars’ gravity well while keeping the Martian moons an accessible target. Some studies have shown the effectiveness of intermediate staging locations such as the aforementioned Mars Base Camp to facilitate missions in the vicinity of Mars and its moons [10,11]. Thus, the establishment of an infrastructure capable of performing In-Situ Resource Utilization (ISRU) of the material found in the regolith and under the surface of Phobos would prove to increase human and robotic exploration capabilities on Mars and its vicinity.
For the purpose of this research, we assumed the existence of a Phobos Base capable of resupplying Mars Base Camp located in a Distant Retrograde Orbit (DRO). We analyzed the necessary trajectories required to transport material from the surface of Phobos to Mars Base Camp. We applied trajectory design, optimization, and control to an orbital transfer in the Circular Restricted Three-Body Problem (CR3BP), specifically from an MP DRO to the surface of Phobos [12]. Given an MP DRO and a landing location on Phobos, the goal is to find the trajectory for which total Δ v for a direct two-burn maneuver is minimum [9,13]. Here, the total Δ v corresponds to the sum of Δ v s required to initiate the transfer in the MP DRO and complete the transfer, i.e., land on the surface of Phobos. Considerations on Time-Of-Flight (TOF) were made in order to ensure that any necessary phasing and/or repositioning maneuvers are taken into account. A 2D analysis is sufficiently accurate since we considered the Mars–Phobos orbit plane as the reference frame, and motion will mostly happen in such a plane. However, further analysis would take into account a three-dimensional space problem in which the gravitational potential of Phobos is also considered [14]. Furthermore, the initial position among the departing DRO will be implemented as the third optimization parameter, given that the minimum Δ v does not necessarily occur at a fixed position along the orbit for a given landing location.
In this paper, a Particle Swarm Optimization (PSO) is used to determine the initial MP DRO given two parameters, [ A x , V y ] . Once the initial DRO is found, a trajectory optimization through a PSO is performed in order to determine the lowest Δ v trajectory to land a specific location on Phobos’ surface, given three parameters to optimize [ α , ϕ , Δ v ] . Here, we also discuss which PSO parameters are used for the analysis and which ones lead to the most efficient (i.e., the fastest) computations, that can be used for other optimization problems in the CR3BP. For further details on such parameters, refer to the nomenclature session at the beginning of the paper or to the detailed Section 2.3 and Section 3.
Phobos is the largest of two moons orbiting Mars, with an approximate size of 13.4 km × 11.2 km × 9.2 km [15,16]. Phobos has an unusual shape (somewhat similar to a potato), which creates numerous difficulties for spacecraft to orbit and therefore land on the moon itself. In the literature, the gravitational potential and the density of Phobos have been of high interest, since it becomes essential for a precise and complete analysis to acknowledge such characteristics, especially for long-duration stability of orbits in the vicinity of the Moon [17,18,19].
The complex gravitational potential and density distribution are the biggest challenges when it comes to orbiting and landing on and from the surface of Phobos. Various methods have been developed to calculate the gravitational potential around irregular-shaped asteroids, highlighting the efforts to overcome such challenges, such as the mascons method [20,21]. In such a method, the celestial body’s shape is approximated by a set of point masses. Other approaches evaluate the gravitational potential of a homogeneous polyhedron, employing a combination of planar triangles on its surface, to describe the gravitational potential of an unusual-shaped body [22,23,24]. This method is considered particularly effective in describing the gravitational field near or on the surface of a constant-density polyhedron [25,26]. As demonstrated in [27,28], the mascon gravity framework can be used with a polyhedral-shaped source. These studies have shown the effectiveness of modeling the gravitational field by calculating the mass of each tetrahedron in the polyhedral shape and assigning it to a point mass at the center of the tetrahedron. This method provides a more accurate estimation of the potential compared to harmonic coefficients and significantly reduces computational processing time [29]. Note that the polyhedral shape of Phobos is available in the Small Bodies Data Ferret [30]. A brief explanation of how the gravitational potential of Phobos is taken into account in this analysis can be found in Section 3.
In this paper, the following approximations are made. Phobos orbits Mars on an elliptical orbit with an eccentricity of e = 0.0151 . For this study, we modeled Mars and Phobos using the CR3BP, which assumes that the eccentricity is 0. The semi-major axis of this orbit (i.e., average distance between Mars and Phobos) is 9376 km. Finally, the orbit’s inclination with respect to Mars’ equatorial plane is around 1°. Since Phobos’ Sphere of Influence (SOI) is below its physical surface, it becomes impossible to orbit the moon in the classical Keplerian sense. In the CR3BP, considering Phobos and Mars as the two primary masses, some families of periodic orbits exist, such as Halo Orbits or Lyapunov Orbits, as shown in Figure 1. However, such periodic orbits are either highly unstable or come dangerously close to the surface [31]. That being considered, the family of DROs is of use in the following analysis. Such family of orbits proved to be “far” enough from Phobos’ surface and to be sufficiently stable as well, up to multiple hundreds of years, when considering a full-force model [31]. The word “far” is here used loosely, since these orbits are within tens of hundreds of km from Phobos. In fact, the orbit is “distant” in the sense that it passes above the Lagrange points, rather than being near the moon, and it is “retrograde” because the spacecraft travels around the moon in the opposite direction that the moon travels around the planet.
Several factors play an essential role while landing a celestial body, such as the Solar Radiation Pressure (SRP), the gravitational influence of the Sun, and the eventual interactions with asteroids approaching the Mars–Phobos system. However, additional information is required regarding the geometry and materials of the spacecraft to perform a complete analysis of how the SRP will affect landing trajectories. Without such information, SRP cannot be taken into account. Since this study only focuses on trajectory optimization and control rather than mission planning, this factor is not considered. Moreover, the effect of the Sun’s gravity becomes relevant in a time span of many months or years. The maneuvers here described have a time of flight of some hours, which is not enough to have the Sun’s gravity play a significant role in the analysis. Finally, it would be virtually impossible to predict asteroids’ flybys and approaches without enormously increasing the computational complexity of the analysis. In the event of an asteroid approaching, this would likely result in a quick passage in the vicinity of Mars and would hardly have any influence on both the DRO stability and the landing maneuvers.
Mission analysis for Earth to Mars–Phobos DROs is of high interest when associated with human and robotic exploration [4,10,11,16]. Phobos could prove extremely useful for future Mars missions as a popular location for future ISRU plants. It is widely recognized that having such in-orbit and in situ infrastructures would make space exploration much more affordable, sustainable, and effective [16,32,33]. Several advantages of such stations include the ability to not have to bring all the necessary material from Earth, including propellant, for deep space missions [34]. Such projects are already in development (such as NASA’s Artemis missions [35]) for what concerns Earth’s Moon: building a refueling in-orbit station along with several on-the-surface infrastructures can easily be imagined as an intermediate step to what would be a deep space exploration mission.
In the following analysis, optimization algorithms are developed in order to generate a specific DRO given a desired distance from Phobos and optimize a landing maneuver to a specific landing site on the surface of Phobos with respect to the Δ v needed, i.e., minimize the necessary impulsive maneuver applied for landing.

2. Background Theory

For the purpose of giving a better insight of the problem, a brief overview of the basic aspects regarding Distant Retrograde Orbits and the Particle Swarm Optimization method are given in the following sections.

2.1. Distant Retrograde Orbits

Distant Retrograde Orbits (DROs) are large orbits around the smaller primary body in the 3BP. Perfectly periodic DROs exist in the CR3BP and in-plane velocity perturbations create quasi-periodic orbits. Such orbits are favorable when looking for quarantine orbits since their stability demonstrates a noteworthy capability to resist perturbations thanks to their interactions with two Lagrange points (L 1 and L 2 ) of the planet–moon system [33]. Indeed, DROs can stay quasi-periodic when larger perturbations than other families of three-body orbits can withstand are applied [31]. On the other hand, both departing and entering a DRO require a significantly high Δ v , since they are marginally stable. Additionally, a spacecraft in a DRO has a longer orbital period and a slower velocity compared to closer orbits. This makes docking and rendezvous challenging since a spacecraft missing its departure/arrival window would need to wait one full orbital period before another opportunity becomes available. Although DROs are highly stable, they are not entirely impervious to external influences. Significant perturbations or disturbances, such as variations in gravitational forces from other celestial bodies or non-uniform mass distributions, can impact the stability and longevity of DROs. However, their relative stability is demonstrated by Wallace et al. [31] using full-force ephemeris models. DROs have been recently proposed to be the ideal family of orbits when it comes to locating in-space infrastructures, and this is due to their outstanding stability and ease of access in terms of a gravity well [33]. DROs are usually characterized by the so-called x-amplitude ( A x ), which represents the largest distance from m 2 in the CR3BP, i.e., Phobos, in the x-axis direction, given a x y z rotating reference frame orbiting the primaries, as shown in Figure 2. In such analysis, it is typical to use a frame of coordinates with its center located at the barycenter of the primary masses [16]. According to this, several essential parameters can be determined in order to define a DRO. Such parameters are the radius at periapsis, and its corresponding velocity, required at the intersection of the x z -plane (i.e., where y = 0 ) in order to create a periodic orbit. Since the radius has only one component along the x-axis, which is A x -dependent, we can say that the essential parameters necessary to define a DRO are [ A x , V y ] , where V y indicates the y-component of the velocity vector at such location, which is, in fact, the only non-zero component; the x-component and z-component of the velocity vector at the considered starting location are zero. DROs have been identified and studied in the literature, for both Earth–Moon DROs and Mars–Phobos DROs [16,36], as shown in Figure 3. However, in this study, a PSO algorithm will be used to determine a DRO given its x-amplitude, which will be the starting point for the landing trajectory optimization.

2.2. Particle Swarm Optimization

The Particle Swarm Optimization method is used in this study as an optimization algorithm to generate periodic DROs and compute Δ v -optimal landing trajectories. PSO is a heuristic method widely used in previous astrodynamics problems [37,38,39]. It is based on the unpredictable motion of bird flocks while searching for food, taking advantage of the mechanism of information sharing that affects the overall behavior of a swarm [40,41]. The initial population that composes a swarm is randomly generated at the first iteration of the algorithm. Each particle is associated with a position vector and a velocity vector. In the PSO terminology, the words “position” and “velocity” of a particle are not to be interpreted in the classical sense (i.e., position and velocity as intended in classical physics and mechanics). Instead, they indicate the position vector containing the unknown parameters to optimize and the velocity vector determining the position update. Each particle represents a possible solution to the problem. At the end of the iterations, the particle corresponding to the optimal solution is selected, i.e., the particle that minimizes the parameters of interest, such as Δ v . At each iteration, the position vector’s elements move in the velocity vector’s corresponding direction, to a new position, which represents a new possible solution. The updating of position and velocity is based on the objective function evaluation at the end of every iteration. The objective function (henceforth referred to as “cost function”) is an expression that needs to be minimized (or maximized). Once this condition is satisfied or the maximum number of iterations set N max is reached, the algorithm will have obtained an optimal solution. At every iteration, the best cost function is used to determine how to update both velocity and position. Indeed, with the approach formulated above, the algorithm will converge on the best solution in every iteration. For each particle, the formula for velocity update includes three components with stochastic weights. In the PSO terminology, these are known as the inertial ( c I ), cognitive ( c C ), and social ( c S ) components. The inertial parameter is usually chosen randomly, but it has appeared to be proportional to each particle’s velocity in the previous iteration for some applications [39]; the cognitive parameter is based on the best position experienced by the particle; finally, the social component is direct toward the personal best position (i.e., the best location yet located by any particle in the swarm). The algorithm terminates when the maximum number of iterations (user decision) is reached. For this study, the following values are used in the algorithm. The function rand stands for a single uniformly distributed random number in the interval (0,1). Adopting the settings suggested in [42,43,44], the inertial, cognitive, and social (stochastic) weights have the following expressions:
c I = ( 1 + r a n d ) 2 c C = 1.49445 · r a n d c S = 1.49445 · r a n d
The constants in Equation (1) are derived from the literature and optimized for this problem in space trajectories [45,46]. The implementation of the PSO was run in MatLab where the function unifrnd was used to assign the initial values at each particle element for each particle. The function unifrnd creates a uniformly distributed array of random values included in a specific range as given by the boundary values in Table 1. Once that is implemented, the initial position is assigned
r 0 = [ A x + ( 1 λ ) ] i ^ + 0 j ^
where
λ = m Phobos ( m Mars + m Phobos )
represents the mass ratio. Note that, as already mentioned, the following analysis will be conducted assuming 2D motion. Therefore, the i th particle (i.e., the particle corresponding to the j th iteration) is
P i = [ V y ( i ) ; T ( i ) ]
so that
v 0 = 0 i ^ + P i ( 1 ) j ^
The PSO algorithm here used is summarized in the flowchart in Figure 4. The coordinate system used is represented in Figure 5. For a matter of simplicity, in Figure 5, m 1 = m Mars and m 2 = m Phobos as they are indicated in Equation (3). Note that these two parameters are picked for the MP DRO analysis described in the Mars–Phobos DRO section below and are here used to better describe the general problem and give the reader a full comprehension of the general approach. Once the Initial Conditions (ICs) are determined ( r 0 and v 0 ), numerical integration is performed in order to compute the resulting trajectory using the CR3BP equations, as derived in [47]:
x ¨ 2 y ˙ x = ( 1 λ ) ( x + λ ) r 1 3 λ ( x 1 + λ ) r 2 3 y ¨ + 2 x ˙ y = ( 1 λ ) y r 1 3 λ y r 2 3 z ¨ = ( 1 λ ) z r 1 3 λ z r 2 3
r 1 = ( x + λ ) 2 + y 2 + z 2 r 2 = ( x + λ 1 ) 2 + y 2 + z 2
Once the integration is finished, if the final position and velocity of the resulting trajectory are within a small tolerance (here 10 10 ) to the initial conditions r 0 and v 0 , then a periodic orbit is achieved and the optimization algorithm has reached a final solution.
Therefore, the cost function, i.e., function to minimize, is
J = r final r 0 + v final v 0 = r final i ^ [ A x + ( 1 λ ) ] + v final j ^ V y ( i )
Satisfying the condition in Equation (8) results in having a periodic orbit, i.e., r ( 0 ) = r ( t f ) and v ( 0 ) = v ( t f ) . Once the cost function is evaluated, each particle velocity (intended in the PSO terminology) is updated as follows:
V i = c I V i + c C ( P best P i ) + c S ( G best P i )
where P best is the personal best of each particle, and G best is the overall best position between all particles and all iterations as of the i th iteration. The coefficients c S , c C , c I are defined in Equation (1). Note that the subscript i in the Equations (4), (8) and (9) indicates the i th element at the j th iteration of the algorithm. Naturally, the larger the number of particles initialized, the more probability the algorithm has to achieve an optimal result. Similarly, a larger number of maximum iterations has more probability to converge to a solution. Nevertheless, a compromise between computational complexity and the number of iterations and particles needs to be made. In fact, increasing either the number of particles or the maximum iterations allowed can make the computational time increase significantly. However, due to the random nature of the algorithm, one cannot guarantee that an absolute optimal solution can be reached. On the other hand, a ‘good’ initial guess is not necessary in order to initiate the algorithm, unlike gradient-based methods. Furthermore, heuristic optimization methods are capable of finding absolute maxima/minima regardless of ICs [45]. On the other hand, the heuristic method of a PSO does not guarantee that a final solution will be found since finding absolute maxima/minima is not guaranteed. Hence, the final solution is sub-optimal.

2.3. Mars–Phobos DRO

Several DROs have been defined in the literature [12,16] for given A x values. Here, a PSO algorithm has been implemented to generate DROs. The dimensionality of the trade space needs to be minimized to optimize the computational effort of the analysis. Since a 2D analysis is performed, considering the state vector [ x , y , z , x ˙ , y ˙ , z ˙ ] , all the z-components are zero, along with the y-component and the x ˙ -component. Hence, two components are left to be found, knowing the initial value for one of those ( y ˙ ). This reduces the minimum parameters necessary to optimize and characterize a DRO to 2. Hence, the optimization was implemented starting from the two parameters [T, V y ], where V y was defined in Section 2 and T represents the orbital period of the DRO. The initial data are summarized in Table 2.
Starting from data taken from [16], the value of V y for A x = 100 km is known to be V y = 0.045620256764708 km/s, which identifies a DRO with an orbital period T = 2.731044880670166 × 10 4 s [48]. Knowing such parameters helps the implementation of boundary values for the two parameters used in this analysis (i.e., the maximum and minimum values that the parameters can assume in the particles). In particular, the initial conditions for the optimization are listed in Table 1. Here, non-dimensionalization parameters were initialized to implement boundary values for the parameters used. The characteristic length l * and the characteristic time t * were implemented as
l * = R = 9376   km
t * = R 3 ( μ Phobos + μ Mars )
Hence, the entire analysis was conducted using non-dimensional values. Similarly, any other DRO of interest can be derived, although only one is here chosen for the results shown.
The parameters to optimize are the initial velocity v 0 (Equation (5)) at position r 0 (Equation (2)) and the orbital period T (Equation (4)). In the literature, different DROs have been identified for A x [ 15 , 300 ] km [16]. DROs with A x values below 15 km result in a certain impact with Phobos’ surface and the ones with higher A x values than 300 km are considered too far away to maintain the characteristics in which we are interested (vicinity to Phobos and low- Δ v for landing). Given the initial orbital parameters for Mars–Phobos and the initial values for the DRO, specified in Table 2, the resulting DRO is shown in Figure 6, and its reference values for position and velocity are summarized in Equation (12).
r = [ 9500.920971466154 , 0.006046323363597 ] km v = [ 0.0007100917133235669 , 0.056800560873732 ] km / s T = 27,310.98049729574 s
where T represents the total orbital period, and v is the velocity vector at position r. Given the N max and N particles specified in Table 1, the result was achieved with a minimum cost function J in the order of 10 4 approximately, as shown in Figure 7, so the final values are such that r f r 0 0.1 mm and v f v 0 0.1 mm/s. The total average computation time was ∼20 s. Several other approaches were attempted, in terms of the number of particles and the number of maximum iterations, but these parameters proved to give a sufficiently accurate result in a reasonable amount of time. If, for instance, we doubled the N max up to 160, the computation time increases up to ∼40 s, generating a minimum cost function J in the order of 10 5 . So doubling the computational time will generally give a result that is just one order of magnitude more accurate than the initial one. As already outlined, one has to compromise between computational efficiency, computational speed, and the results’ accuracy.
Many computational issues may occur while implementing the algorithm. However, parallel programming and computing is a suitable solution to speed up the computation time. In fact, every possible trajectory computed at each iteration is independent with respect to all the other trajectories, until every particle is updated. Note that for this study a 2016 MacBook Pro is used. Hardware and Software details are summarized in Table 3.

3. Landing Trajectory Optimization

In this section, the trajectory optimization for multiple landing maneuvers is analyzed through the use of PSO. Given an initial stable parking DRO around Phobos like the example computed in Section 2.3, and a desired landing location, the analysis aims to find the minimum Δ v necessary to encounter that specific location. Here, the gravitational parameters J 2 and J 3 of Phobos are taken into consideration during the CR3BP equations of motion integration. Although a simplified geometry of Phobos is considered, taking the gravitational distribution into account gives more accurate results to the presented analysis. The problem is computationally more complex than the one described in the MP DRO Section and it presents several more difficulties in the formulation. First of all, the choice of the parameters to optimize is non-trivial and there are different ways to approach the optimization. Here, the following parameters were chosen, minimizing the dimensionality of the trade space to three, as explained in Section 2.3:
P i = [ α ( i ) ; Δ v ( i ) ; ϕ ( i ) ]
where the element Δ v ( i ) represents the Δ v to apply, and α ( i ) indicates the angle with respect to the x-axis at which the Δ v is applied. The parameter ϕ ( i ) represents the angle with respect to the x-axis, which consequently identifies a location on the DRO r ( ϕ ) , as shown in Figure 8. In this figure, a graphical representation of the optimization parameters is shown. As represented, ϕ indicates the position of the spacecraft along the initial DRO (we can compare ϕ to a sort of true anomaly). Once the location of the spacecraft is known, the entire space vector of the spacecraft at such location is also known. Here, a Δ v needs to be applied to depart the DRO and initiate the landing maneuver. Hence, the other parameters to optimize are the Δ v magnitude and the angle α at which the impulsive maneuver is applied (no second out-of-plane angle is necessary since this is purely a 2D problem). This α angle is essential to calculate the resulting trajectory and perform a vector analysis. Note that in this analysis, the index i represents the i th particle at the j th iteration.
This analysis requires more computational time than the DRO optimization because there are three parameters to optimize instead of two, and we do not have specific analytic expressions to relate them to each other. Hence, a larger number of iterations was performed and a larger amount of particles was initialized, increasing the computational time and the overall complexity of the analysis. Computational times for the landing trajectory optimization are reported in Table 4. The starting DRO is equal to the one defined in the optimization described in Section 2.3. Therefore, as obtained in Equation (12), the initial values and the particle boundary values are summarized in Table 5 and Table 6, respectively. As explained in Section 2.3, the dimensionality of the trade space is here minimized, and only three parameters are necessary to define a landing trajectory since most of the components of the state vector are either known or null.
In this particular optimization, a penalty scaling factor γ was introduced. The addition of such a factor is necessary in order to prioritize the terms of the cost function [49]. In fact, the essential condition for a valid outcome of the optimization is to land at the desired location on Phobos. Indeed, a trajectory is considered valid only if it intersects the surface of Phobos at the desired coordinates. Secondarily, given that a valid trajectory to accomplish such a condition has been found, the total Δ v needs to be minimized. Hence, a penalty scaling factor plays an essential role in prioritizing the parameters and improving the algorithm’s reliability to obtain a close-to-optimal solution. The expression for γ was derived by Conte et al. [49] and it is designed to be an adaptive non-linearly decreasing function of the iterations number i [50]. The following proved to be an effective value for this factor:
γ ( i ) = β γ initial γ initial γ final N max ( 2 N max i ) i
where i is the current iteration number, N max is the maximum number of iterations and γ initial and γ final represent the initial and the final weighting values. In this analysis, they are equal to 1 and 0.1, respectively. The factor β is an order scaling factor, set equal to 100 for this optimization, which is used to make sure that the location checking term has a higher order than the Δ v term. Therefore, the cost function becomes
J = γ r final r landing + Δ v total , if r final r landing > tol Δ v total , if r final r landing < tol
For this optimization, the tolerance (tol) is set to 0.1 km (100 m). It is clear how the scaling factor prioritizes a specific term in Equation (15), given that the Δ v term is a relatively small value.
For a matter of simplicity, Phobos was ideally divided into 12 sectors, and a landing location within each sector was picked as a reference location for such an area. This means that the Δ v required to achieve landing in such a sector can be considered quantitatively similar to the reference location’s Δ v . Results for the landing locations in each sector are summarized in Table 7. Computation times for each example in Table 7 are indicated in Table 4. As expected, the average computation time is slightly higher than the one for the DRO optimization described in the MP DRO analysis (which is in a range between ∼30 s and ∼1 min).
The landing trajectories for each sector are represented in Figure 9 and a heat map of all the landing locations is shown in Figure 10. To have a better insight into the results, a fully loaded SpaceX Starship is considered. SpaceX’s Starship is expected to have an empty mass of ∼120 t (260,000 lb or 120,000 kg) and a gross mass of ∼1320 t (2,910,000 lb or 1,320,000 kg) with a propellant capacity of ∼1200 t (2,600,000 lb or 1,200,000 kg) [51]. Recall Tsiolkovsky’s rocket equation:
Δ v = v e ln m 0 m f = g 0 I s p ln m 0 m f
where m 0 is the total mass of the spacecraft, including the propellant, m f is the final total mass assuming all propellant to be consumed (also known as dry mass), I s p is the spacecraft’s specific impulse and g 0 is the standard gravity at Earth’s sea level. Note that the SI-metric ton (t) is used, where 1 t = 1000 kg ≃ 2205 lb.
We can now assume to land on Phobos’ surface at sector 7, the latitude of which corresponds to the Stickney Crater on the surface of Phobos and which corresponds to a total Δ v of 27.41 m/s (see Table 7). From Equation (16), we can estimate the necessary propellant mass to land the full-load Starship on Phobos.
We consider the spacecraft to empty its tanks at landing. Given a specific impulse of 380 s in a vacuum and a payload of 100 t (100,000 kg), Starship will achieve landing with a total propellant mass equal to
m p = m f ( e Δ v g 0 I s p 1 ) = ( 120 + 100 ) · ( e 27.41 9.81 · 380 1 ) 1.62 t 1620 kg
m p m f = 0.7 %
where m p is the total propellant mass used to achieve the indicated Δ v . Since the landing Δ v is quantitatively equal to the Δ v required to go from the surface of Phobos to the initial DRO (the trajectory will just look different from a qualitative point of view), and we assumed Phobos to have a re-fueling station on its surface, the total propellant mass necessary to leave and arrive at the DRO is simply two times the mass indicated in Equation (17). Therefore, assuming one travel per terrestrial day for a year (365 days) back and forth the Stickney Crater, the total propellant mass per year necessary to achieve such transfers is ∼1185.22 t of propellant, which is equal to ∼1,185,220 kg.
If we consider a 1200 t propellant capacity, Starship can achieve a back-and-forth transfer ∼369 times before refueling. Although it is challenging to precisely estimate the propellant price per ton, SpaceX claims that refueling Starship would cost approximately USD 900,000. This means that a complete transfer would cost around USD 2435, transporting a 100 t payload.

4. Conclusions

This study led to valuable results for both DRO implementations and landing trajectory optimization, proving the feasibility of landing trajectories to the surface of Phobos. Several DROs have been defined in the literature for given A x values, but not much exists about landing trajectories from DROs to the surface of Phobos. Regarding the landing trajectory optimization, which is the main purpose of this study, it is clear from Table 7 how the Δ v is on average equal to ∼30 m/s. This value is reasonably low and makes this orbit feasible for an in-orbit refueling station or in general for a stationary orbit around Phobos, which one can easily reach and leave. Several landing locations were reached, and the impulsive Δ v was always found to be in approximately the same range. Therefore, nearly every point on Phobos’ surface can be reached with a Δ v of approximately ∼30 m/s. Furthermore, as summarized in Table 4, the average computation time for such a PSO algorithm is reasonable and makes the optimization feasible for future mission designs.
Mars’ moons exploration has been of high interest within the scientific community in the past years. The existence of periodic orbits around Phobos, such as DROs, makes the future design of in-orbit refueling and supporting stations more than feasible. Moreover, the ease of departing and arriving in such orbits with low Δ v s gives the possibility of exploring Phobos and establishing a Mars Base Camp, in order to explore Mars and its surroundings with great versatility and with lower flight times between transfers from and to Mars itself. Having stations or spacecrafts orbiting around Mars (in Low-Mars Orbits for example) would require continuous trajectory adjustments because of Mars’ gravity potential acting on the spacecraft/station, and will then result in significant Δ v used over the course of a long-term mission. Besides, given the distance between Earth and Mars, it would not be convenient to perform frequent trajectory adjustments or remote control operations, given an average transmission delay of ∼20 min between Mars and Earth. Hence, having a long-term stability solution is crucial, and DROs appear to be feasible for such an objective. The method here presented can be used for multiple Mars–Phobos DROs optimizations at different amplitudes. Plus, a Mars–Deimos DRO can utilize this very same approach.

Author Contributions

V.B. is the primary author and researcher of this paper who developed the mathematical tools and performed all major calculations presented in the paper. D.C. is Vittorio’s faculty advisor, who helped guide Vittorio’s research work and reviewed the content and writing of this paper through multiple rounds of reviews. Conceptualization, V.B. and D.C.; formal analysis, V.B.; methodology, V.B.; project administration, D.C.; supervision, D.C.; writing—original draft, Vittorio Baraldi; writing—review and editing, D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Δ v Change in velocity (km/s)
eOrbital eccentricity
A x Larger x-coordinate from the primary body in the Circular Restricted
Three-Body Problem (km)
V y y-component of velocity (km/s)
N max Maximum number of iterations used in Particle Swarm
Optimization (PSO)
TOrbital Period (s)
P i Number of particles used in PSO
JCost function
λ Mass ratio of Mars and Phobos
r Position vector
v Velocity vector
N particles Number of particles in PSO
B lo Lower limit for particles in PSO
B up Upper limit for particles in PSO
α Angle with respect to the x-axis at which the Δ v is applied (°)
ϕ Position along the orbit with respect to the x-axis, starting at r 0 (°)
r 0 Initial position vector (km)
v 0 Initial velocity vector (km/s)
r final Final position vector (km)
v final Final velocity vector (km/s)
γ Penalty scaling factor used by PSO
μ Gravitational parameter (km 3 /s 2 )
m 0 Total spacecraft mass (t)
m f Dry spacecraft mass (t)
m p Propellant mass (t)
I s p Specific Impulse (s)
g 0 Standard acceleration due to gravity at Earth’s sea level (m/s 2 )
v e Exhaust velocity (m/s)
PSOParticle Swarm Optimization
DRODistant Retrograde Orbits
CR3BPCircular-Restricted Three-Body Problem

References

  1. Loff, S. Lunar Gateway. 2012. Available online: https://www.nasa.gov/topics/moon-to-mars/lunar-outpost (accessed on 12 April 2020).
  2. Johnson, S.K.; Mortensen, D.J.; Chavez, M.A.; Woodland, C.L. Gateway—A communications platform for lunar exploration. In Proceedings of the 38th International Communications Satellite Systems Conference (ICSSC 2021), Arlington, VA, USA, 27–30 September 2021; Volume 2021, pp. 9–16. [Google Scholar] [CrossRef]
  3. National Aeronautics and Space Administration (NASA). White Paper: Gateway Destination Orbit Model: A Continuous 15 Year NRHO Reference Trajectory; NASA: Washington, DC, USA, 2019. [Google Scholar]
  4. JAXA. Martian Moons Exploration: Mission Overview. 2017. Available online: http://www.isas.jaxa.jp/en/topics/files/MMX170412_EN.pdf (accessed on 8 June 2023).
  5. Martin, L. Mars Base Camp. Available online: https://www.lockheedmartin.com/en-us/products/mars-base-camp.html (accessed on 8 June 2023).
  6. Ulamec, S.; Michel, P.; Grott, M.; Boettger, U.; Hübers, H.W.; Schröder, S.; Cho, Y.; Rull, F.; Murdoch, N.; Vernazza, P.; et al. Surface science on Phobos with the MMX Rover. In Proceedings of the 44th COSPAR Scientific Assembly, Athens, Greece, 16–24 July 2022. [Google Scholar]
  7. Adams, E.; Murchie, S.; Eng, D.; Chabot, N.; Guo, Y.; Arvidson, R.; Trebi-ollennu, A.; Seelos, F. Mission concept for robotic exploration of Deimos. In Proceedings of the 61st International Astronautical Congress 2010 (IAC 2010), Prague, Czech Republic, 27 September–1 October 2010; Volume 11, pp. 9249–9259. [Google Scholar]
  8. Ulamec, S.; Michel, P.; Grott, M.; Boettger, U.; Hübers, H.W.; Cho, Y.; Rull, F.; Murdoch, N.; Vernazza, P.; Biele, J.; et al. The MMX Rover Mission to Phobos: Science Objectives. In Proceedings of the 72nd International Astronautical Congress (IAC 2021), Dubai, United Arab Emirates, 25–29 October 2021. [Google Scholar]
  9. Hopkins, J.; Pratt, W. Comparison of Phobos and Deimos as Destinations for Human Exploration, and Identification of Preferred Landing Sites. In Proceedings of the AIAA Space 2011 Conference and Exposition, Long Beach, CA, USA, 27–29 September 2011. [Google Scholar]
  10. Cichan, T.; Norris, S.; Chambers, R.; Jolly, S.; Scott, A.; Bailey, S. Mars Base Camp: A Martian Moon Human Exploration Architecture. In Proceedings of the AIAA Space 2016 Conference and Exposition, Long Beach, CA, USA, 13–16 September 2016. [Google Scholar]
  11. Chican, T.; Bailey, S.; Antonelli, T.; Jolly, S.; Chambers, R.; Ramm, S. Mars Base Camp: An Architecture for Sending Humans to Mars. New Space 2017, 5, 203–218. [Google Scholar]
  12. Conte, D. Determination of Optimal Earth-Mars Trajectories to Target the Moons of Mars; The Pennsylvania State University (Department of Aerospace Engineering): State College, PA, USA, 2014. [Google Scholar]
  13. Bezrouk, C.; Parker, J. Ballistic Capture into DROs around Phobos: An Approach to Entering Orbit around Phobos without a Critical Maneuver near the Moon. Celest. Mech. Dyn. Astron. 2018, 130, 10. [Google Scholar] [CrossRef]
  14. Fraeman, A. Analysis of Disk-Resolved OMEGA and CRISM Spectral Observation of Phobos and Deimos. J. Geophys. Res. Planets 2012, 117, 1991–2012. [Google Scholar] [CrossRef] [Green Version]
  15. Wessen, A. Phobos: By the Numbers. Available online: http://solarsystem.nasa.gov/moons/mars-moons/phobos/by-the-numbers/ (accessed on 15 March 2020).
  16. Conte, D.; Spencer, D. Mission Analysis for Earth to Mars-Phobos Distant Retrograde Orbits. Acta Astronaut. 2018, 151, 761–771. [Google Scholar] [CrossRef]
  17. Borderies, N.; Yoder, C. Phobos Gravity Field and its Influence on its Orbit and Physical Libration. Astron. Astrophys. 1990, 233, 235–251. [Google Scholar]
  18. Shi, X.; Willner, K.; Oberst, J.; Ping, J.; Ye, S. Working models for the gravitational field of Phobos. Sci. China Phys. Mech. Astron. 2012, 55, 358–364. [Google Scholar] [CrossRef]
  19. Maistre, S.L.; Rivoldini, A.; Rosenblatt, P. Signature of Phobos’ Interior Structure in its Gravity Field and Libration. Icarus 2019, 321, 272–290. [Google Scholar] [CrossRef]
  20. Muller, P.M.; Sjogren, W.L. Mascons: Lunar Mass Concentrations. Science 1968, 161, 680–684. [Google Scholar] [CrossRef]
  21. Muller, P.; Sjogren, W. Consistency of Lunar Orbiter residuals with trajectory and local gravity effects. J. Spacecr. Rocket. 1969, 6, 849. [Google Scholar] [CrossRef]
  22. Werner, R. The gravitational potential of a homogeneous polyhedron or don’t cut corners. Celest. Mech. Dyn. Astron. 1994, 59, 253–278. [Google Scholar] [CrossRef]
  23. Werner, R.; Scheeres, D. Exterior Gravitation of a Polyhedron Derived and Compared with Harmonic and Mascon Gravitation Representations of Asteroid 4769 Castalia. Celest. Mech. Dyn. Astron. 1996, 65, 313–344. [Google Scholar] [CrossRef]
  24. Werner, R. Spherical harmonic coefficients for the potential of a constant-density polyhedron. Comput. Geosci. 1997, 23, 1071–1077. [Google Scholar] [CrossRef]
  25. Scheeres, D.; Ostro, S.; Hudson, S.; Dejong, E.; Suzuki, S. Dynamics of Orbits Close to Asteroid 4179 Toutatis. Icarus 1998, 132, 53–79. [Google Scholar] [CrossRef] [Green Version]
  26. Scheeres, D.; Williams, B.; Miller, J. Evaluation of the Dynamic Environment of an Asteroid: Applications to 433 Eros. J. Guid. Control Dyn. 2023, 23, 466. [Google Scholar] [CrossRef]
  27. Chanut, T.; Aljbaae, S.; Carruba, V. Mascon gravitation model using a shaped polyhedral source. Mon. Not. R. Astron. Soc. 2015, 450, 3742–3749. [Google Scholar] [CrossRef]
  28. Aljbaae, S.; Chanut, T.; Carruba, V.; Souchay, J.; Prado, A.; Amarante, A. The dynamical environment of asteroid 21 Lutetia according to different internal models. Mon. Not. R. Astron. Soc. 2016, 464, 3552–3560. [Google Scholar] [CrossRef] [Green Version]
  29. Aljbaae, S.; Souchay, J.; Carruba, V.; Sanchez, D.; Prado, A. Influence of Apophis’ Spin Axis Variations on a Spacecraft during the 2029 close Approach with Earth. Rom. Astron. J. 2022, 1, 317–338. [Google Scholar]
  30. NASA. Small Bodies Data Ferret. Available online: https://sbnapps.psi.edu/ferret/ (accessed on 15 July 2023).
  31. Wallace, M.; Parker, J.; Strange, N.; Grebow, D. Orbital Operations for Phobos and Deimos Exploration. In Proceedings of the AIAA/AAS Astrodynamics Specialist Conference, Minneapolis, MN, USA, 13–16 August 2012; p. 5067. [Google Scholar]
  32. Scott, C.; Spencer, D. Stability Mapping of Distant Retrograde Orbits and Transports in the Circular Restricted Three-Body Problem. In Proceedings of the AIAA/AAS Astrodynamics Specialist Conference and Exhibit, Honolulu, HI, USA, 18–21 August 2008. [Google Scholar] [CrossRef]
  33. Bezrouk, C.; Parker, J. Long Duration Stability of Distant Retrograde Orbits. In Proceedings of the AIAA/AAS Astrodynamics Specialist Conference, San Diego, CA, USA, 4–7 August 2014. [Google Scholar]
  34. Ho, K.; de Week, O.; Hoffman, J.; Shishko, R. Dynamic Modeling and Optimization for Space Logistics using Time-Expanded Networks. Acta Astronaut. 2014, 105, 428–443. [Google Scholar] [CrossRef]
  35. NASA. Artemis Plan: NASA’s Lunar Exploration Program Overview. 2020. Available online: https://www.nasa.gov/sites/default/files/atoms/files/artemis_plan-20200921.pdf (accessed on 8 June 2023).
  36. Conte, D.; Carlo, M.D.; Ho, K.; Spencer, D.; Vasile, M. Earth-Mars Transfers through Moon Distant Retrograde Orbits. Acta Astronaut. 2018, 143, 372–379. [Google Scholar] [CrossRef] [Green Version]
  37. Pontani, M. Simple Method to Determine Globally Optimal Orbital Transfers. J. Guid. Control Dyn. 2009, 32, 899–914. [Google Scholar] [CrossRef]
  38. Pontani, M.; Ghosh, P.; Conway, B. Particle Swarm Optimization of Multiple-Burn Rendezvous Trajectories. J. Guid. Control Dyn. 2012, 35, 1192–1207. [Google Scholar] [CrossRef]
  39. Pontani, M. Particle Swarm Optimization of Ascent Trajectories of Multi-Stage Launch Vehicles. Acta Astronaut. 2014, 94, 852–864. [Google Scholar] [CrossRef]
  40. Eberhart, R.; Kennedy, J. A New Optimizer using Particle Swarm Theory. In Proceedings of the 6th International Symposium on Micromachine and Human Science, Nagoya, Japan, 4–6 October 1995. [Google Scholar]
  41. Eberhart, R.; Kennedy, J. Particle Swarm Optimization. In Proceedings of the IEEE International Conference on Neural Networks, Perth, Australia, 27 November–1 December 1995. [Google Scholar]
  42. Hu, X.; Eberhart, R. Solving constrained nonlinear optimization problems with particle swarm optimization. Citeseer 2002, 2002, 203–206. [Google Scholar]
  43. Hu, X.; Shi, Y.; Eberhart, R. Recent advances in particle swarm. In Proceedings of the 2004 Congress on Evolutionary Computation (IEEE Cat. No. 04TH8753), Portland, OR, USA, 19–23 June 2004; Volume 1, pp. 90–97. [Google Scholar] [CrossRef]
  44. Hu, X.; Eberhart, R.; Shi, Y. Engineering optimization with particle swarm. In Proceedings of the 2003 IEEE Swarm Intelligence Symposium. SIS’03 (Cat. No. 03EX706), Indianapolis, IN, USA, 26–26 April 2003; pp. 53–57. [Google Scholar] [CrossRef]
  45. Pontani, M.; Conway, B. Particle Swarm Optimization applied to Space Trajectories. J. Guid. Control Dyn. 2010, 33, 1429–1441. [Google Scholar] [CrossRef]
  46. Pontani, M.; Conway, B. Particle Swarm Optimization applied to Impulsive Orbital Transfers. Acta Astronaut. 2012, 74, 141–155. [Google Scholar] [CrossRef]
  47. Conte, D.; Spencer, D. Interplanetary Astrodynamics; Routledge: Abingdon, UK, 2023. [Google Scholar]
  48. Conte, D. Semi-Analytical Solutions for Proximity Operations in the Circular Restricted Three Body Problem; The Pennsylvania State University (Department of Aerospace Engineering): State College, PA, USA, 2019. [Google Scholar]
  49. He, G.; Conte, D.; Melton, R.G.; Spencer, D.B. Fireworks Algorithm applied to Trajectory Design for Earth to Lunar Halo Orbits. J. Spacecr. Rocket. 2020, 57, 235–246. [Google Scholar] [CrossRef]
  50. Zheng, S.; Janecek, A.; Tan, Y. Enhanced Fireworks Algorithm. In Proceedings of the 2013 IEEE Congress on Evolutionary Computation, Cancun, Mexico, 20–23 June 2013; pp. 2069–2077. [Google Scholar] [CrossRef]
  51. SpaceX. Starship-Official SpaceX Starship Page. Available online: https://www.spacex.com/vehicles/starship/ (accessed on 12 April 2021).
Figure 1. Mars–Phobos periodic orbits examples in the Circular Restricted Three Body Problem: L 1 halo orbits (red), L 2 halo orbits (green), L 1 vertical orbits (purple), L 2 vertical orbits (orange), and DROs (blue).
Figure 1. Mars–Phobos periodic orbits examples in the Circular Restricted Three Body Problem: L 1 halo orbits (red), L 2 halo orbits (green), L 1 vertical orbits (purple), L 2 vertical orbits (orange), and DROs (blue).
Universe 09 00348 g001
Figure 2. Graphic representation of parameters [ A x , V y ] to optimize to generate a periodic DRO in the CR3BP rotating reference frame x y z .
Figure 2. Graphic representation of parameters [ A x , V y ] to optimize to generate a periodic DRO in the CR3BP rotating reference frame x y z .
Universe 09 00348 g002
Figure 3. Mars–Phobos DRO samples with Lagrange Points L 1 and L 2 for reference.
Figure 3. Mars–Phobos DRO samples with Lagrange Points L 1 and L 2 for reference.
Universe 09 00348 g003
Figure 4. Flowchart for PSO algorithm.
Figure 4. Flowchart for PSO algorithm.
Universe 09 00348 g004
Figure 5. Geometry of the restricted three-body problem.
Figure 5. Geometry of the restricted three-body problem.
Universe 09 00348 g005
Figure 6. Mars–Phobos DRO for A x = 125 km resulting from the particle swarm optimization.
Figure 6. Mars–Phobos DRO for A x = 125 km resulting from the particle swarm optimization.
Universe 09 00348 g006
Figure 7. Cost Function J compared to iteration numbers for different DRO amplitudes. (a) A x = 100 km; (b) A x = 125 km; (c) A x = 150 km.
Figure 7. Cost Function J compared to iteration numbers for different DRO amplitudes. (a) A x = 100 km; (b) A x = 125 km; (c) A x = 150 km.
Universe 09 00348 g007
Figure 8. Graphic representation of parameters α , ϕ and Δ v . These are the parameters to optimize in the landing trajectory analysis of this study.
Figure 8. Graphic representation of parameters α , ϕ and Δ v . These are the parameters to optimize in the landing trajectory analysis of this study.
Universe 09 00348 g008
Figure 9. Landing trajectories for each sector. Details are summarized in Table 7.
Figure 9. Landing trajectories for each sector. Details are summarized in Table 7.
Universe 09 00348 g009
Figure 10. Phobos’ sector division and qualitative landing trajectories.
Figure 10. Phobos’ sector division and qualitative landing trajectories.
Universe 09 00348 g010
Table 1. Boundary values for DRO optimization.
Table 1. Boundary values for DRO optimization.
VariableValueDescription
B lo ( V y ) 3 2 t * l * V y Lower limit for V y (km/s)
B up ( V y ) t * l * V y Upper limit for V y (km/s)
B lo (T)T/ t * Lower limit for T (s)
B up (T) 2 π Upper limit for T (s)
N max 80Maximum number of iterations
N particles 40Number of particles
Table 2. Orbital parameters for Mars–Phobos and initial values for DRO optimization.
Table 2. Orbital parameters for Mars–Phobos and initial values for DRO optimization.
VariableValueDescription
μ Phobos (km 3 s 2 ) 7.11358812096305 × 10 4 Gravitational parameter for Phobos
μ Mars (km 3 s 2 ) 42,828.375214 Gravitational parameter for Mars
λ 1.660952106463386 × 10 8 Mass ratio
R (km)9376Average distance
Mars-Phobos
A x (km)125DRO Amplitude
Table 3. Computer system characteristics.
Table 3. Computer system characteristics.
ItemDescription
Operating SystemMacOS Monterey Version 12.6.7
Processor3.3 GHz Dual-Core Intel Core i7
RAM16 GB 2133 MHz
MatLab versionR2020b Update 3
System architecture64-bit operating system, x-64-based processor
Table 4. Computation times for landing optimization algorithm.
Table 4. Computation times for landing optimization algorithm.
Landing Location SectorComputation Time
12 min, 56 s
22 min, 21 s
33 min, 20 s
42 min, 17 s
52 min, 42 s
62 min, 15 s
72 min, 56 s
83 min, 17 s
92 min, 21 s
102 min, 19 s
112 min, 53 s
124 min, 8 s
Average2 min, 54 s
Table 5. Initial values for landing trajectory optimization.
Table 5. Initial values for landing trajectory optimization.
VariableValueDescription
r 0 (km) [ 9500.920971466154 , 0.006046323363597 ] Initial position of parking DRO
v 0 (km/s) [ 0.0007100917133235669 , 0.056800560873732 ] Orbit velocity at position r 0
T orbit (s)27,310.98049729574Orbital period
Table 6. Boundary values for landing optimization using PSO.
Table 6. Boundary values for landing optimization using PSO.
VariableValueDescription
B lo ( V y )0Lower limit for Δ v (km/s)
B up ( V y ) V y t * l * Upper limit for Δ v (km/s)
B lo ( α )0Lower limit for α (rad)
B up ( α ) 2 π Upper limit for α (rad)
B lo ( ϕ )0Lower limit for ϕ (rad)
B up ( ϕ ) 2 π Upper limit for ϕ (rad)
N max 200Maximum number of iterations
N particles 100Number of particles
Table 7. Final results for landing optimization at one landing location for each sector.
Table 7. Final results for landing optimization at one landing location for each sector.
Landing Loc. SectorTime of FlightTotal Δ v (m/s) α (°) ϕ (°)
14 h, 41 min27.35246.10147.14
23 h, 51 min29.24205.53108.16
35 h, 10 min31.02347.27256.35
47 h, 33 min32.42101.3289.22
54 h, 54 min27.2788.41337.07
63 h, 29 min320270.49
75 h, 13 min27.41100.380
86 h, 6 min28.72154.3262.75
93 h, 41 min31.930270.84
104 h, 24 min31.34168.8273.85
115 h, 36 min27.66260.27204.36
125 h, 13 min30.25209.12161.04
Average4 h, 59 min29.72 m/s
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Baraldi, V.; Conte, D. Trajectory Optimization and Control Applied to Landing Maneuvers on Phobos from Mars-Phobos Distant Retrograde Orbits. Universe 2023, 9, 348. https://doi.org/10.3390/universe9080348

AMA Style

Baraldi V, Conte D. Trajectory Optimization and Control Applied to Landing Maneuvers on Phobos from Mars-Phobos Distant Retrograde Orbits. Universe. 2023; 9(8):348. https://doi.org/10.3390/universe9080348

Chicago/Turabian Style

Baraldi, Vittorio, and Davide Conte. 2023. "Trajectory Optimization and Control Applied to Landing Maneuvers on Phobos from Mars-Phobos Distant Retrograde Orbits" Universe 9, no. 8: 348. https://doi.org/10.3390/universe9080348

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop