1 - Purdue University



Comparative Assessment of Human Missions to Mars

Damon Landau

Ph.D. Preliminary Exam

How Shall We Go to Mars?

The allure of people traveling to Mars has been the inspiration for numerous mission proposals.1–20 Despite decades of comparative analyses,21–33 the debate continues on how to transport a crew to Mars and return them home safely. Indeed, there are myriad design options when one considers the possible combinations of propulsion technologies, mission architectures, transfer vehicles, and timelines for a mission to Mars. New trajectories (e.g. Aldrin’s cycler) and technology developments (especially from NASA’s in-space propulsion program) provide tremendous leveraging effects for the human exploration of Mars. However, a comparative analysis of the costs and benefits of these mission options is unavailable, and inconsistencies in crew number, vehicle masses, and allowable transfer times cloud previous comparative assessments.

To characterize the effects of these mission trades, I calculate the injected mass to low-Earth orbit (IMLEO) with the condition that the crew, payload, vehicles, propulsion systems, and transfer time of flight (TOF) must be consistent among different mission architectures. I choose IMLEO as the metric because a change in mass is the most direct effect of varying a mission parameter and because IMLEO is strongly correlated to the dollar-cost of a given mission.33–36 In this way, the IMLEO benefit from developing a potential technology is known a priori to help direct the path of Mars exploration. The goal of this research is to determine the best way of sending people to Mars by assembling and assessing the array of mission architectures and available technologies for interplanetary exploration.

Mars exploration architectures are differentiated by the placement of the interplanetary transfer vehicle at Earth or Mars (see Table 1). For example, NASA’s Design Reference Mission places the transfer vehicle into a parking orbit at Mars arrival (dubbed a semi-direct architecture).19, 20 Other ideas include parking orbits at both Earth and Mars (stop-over),37, 38 a flyby at both Earth and Mars (cycler),39–44 a flyby at Earth and parking orbit at Mars (Mars-Earth semi-cycler),45, 46 and flyby at Earth and parking orbit at Mars (Earth-Mars semi-cycler).47 A separate taxi vehicle then ferries the crew between the surface of a planet and the transfer vehicle to complete the crew transfer. In a “direct” architecture the transfer vehicle lands on the surface of both planets, eliminating the need for a taxi vehicle. Often, incorporating a parking orbit or flyby at planetary encounters lowers the IMLEO because the relatively massive transfer vehicle performs less maneuvers.

Table 1 Placement of interplanetary transfer vehicle for different architectures

|Architecture |Earth Encounter |Mars Encounter |Schemata |

| | | | |

|Direct |Surface |Surface | |

| | | | |

|Semi-Direct |Surface |Parking Orbit | |

| | | | |

|Stop-Over |Parking Orbit |Parking Orbit | |

| | | | |

|M-E Semi-Cycler |Flyby |Parking Orbit | |

| | | | |

|E-M Semi-Cycler |Parking Orbit |Flyby | |

| | | | |

|Cycler |Flyby |Flyby | |

The available propulsion technology plays a significant role in determining the IMLEO for a Mars mission. For example, chemical propulsion [e.g. liquid hydrogen with liquid oxygen (LH2/LOX)] has been the workhorse for space exploration, but the higher specific impulse of nuclear thermal rockets (NTR) or nuclear electric propulsion (NEP) can reduce propellant mass. Another option is to make the propellant required for the return trip on Mars, or in-situ propellant production. For example, a feedstock of terrestrial hydrogen may be combined with the carbon dioxide at Mars to produce methane and oxygen, eliminating the need to launch the return propellant from Earth. Also, should it become practical to extract large quantities of water from Martian regolith, LH2/LOX propulsion systems may be used without the need for any propellant feedstock. The atmosphere of Mars has been used to decelerate spacecraft for surface landing and to lower the energy of a parking orbit, but aerocapture, where the spacecraft is decelerated from the interplanetary transfer into a parking orbit, has yet to be attempted. Mission architectures that rely on a parking orbit at Earth or Mars can benefit from aerocapture because the aerobrake mass fraction is usually less than the propellant mass fraction for orbit capture. I plan to also assess the benefits of reusable propulsion systems and transporting propellant as cargo from one planet to the other. Table 2 provides a list of potential technologies along with an approximate technology readiness level (TRL).48

Table 2 Current and near-tern technologies

|Technology |Approximate Readiness Levela|

|Chemical Propulsion |9 |

|Parking Orbit Rendezvous (Earth) |9 |

|Reusable Chemical Propulsion |8 |

|Parking Orbit Rendezvous (Mars) |8 |

|Refuel in Orbit (Earth) |8 |

|Cargo Nuclear Electric Propulsion |7 |

|Refuel in Orbit (Mars) |7 |

|Hyperbolic Rendezvous (Earth) |7 |

|Hyperbolic Rendezvous (Mars) |6 |

|Nuclear Thermal Rocket |6 |

|Aerocapture |6 |

|Transfer Vehicle NEP |5 |

|In-Situ Propellant Production |5 |

|Mars Launch Vehicle NTR |4 |

|Mars Water Excavation |3 |

aFor a definition of technology readiness levels see Ref. 48.

To further explore the design space I vary the mass of the transfer vehicle and taxi for a given crew size. For example, a low-mass transfer vehicle is preferred from a IMLEO standpoint, but may be uncomfortable or even detrimental to the health of the crew. By increasing the vehicle mass per person, additional radiation shielding or artificial gravity may be incorporated to reduce mission risk. Another key factor to crew health is the time the crew must spend in space. To examine the effect on IMLEO from reducing the flight-time, the allowed TOF between Earth and Mars may also be varied. Finally, the change in IMLEO due to varying the mass of Mars cargo (e.g. habitat, power plant, etc.) will also be examined.

Earth-Mars Trajectories

Earth-Mars trajectories with low-energy requirements that also limit the (transfer) time a crew spends in interplanetary space are essential to the design of cost-effective, minimal-risk missions. To examine the effects of limiting TOF, I compute optimal ΔV trajectories (to reduce mission cost) with constrained TOF (to reduce mission risk). Traditionally, Mars trajectories fall into two categories: 1) opposition class49–57 for short duration mission (600 days) with short Mars stay time (30 days), and 2) conjunction class53–63 for long Mars stay time (550 days) with long mission durations (900 days). Unfortunately, trajectories between Earth and Mars with short TOF, short mission duration, and low energy requirements (low ΔV) do not exist; thus, I do not examine opposition class trajectories further. It is assumed that one mission (to Mars and back) occurs once every synodic period. Since Earth-Mars trajectories approximately repeat every seven synodic periods (every 14.95 years), I compute short TOF trajectories over a seven synodic-period cycle (for 2009–2022 Earth departure years). The effects of powered versus aero-assisted planetary arrivals is also examined.64–67

Conjunction class trajectories (which are fairly well understood and documented) are suitable for direct, semi-direct, and stop-over missions because these architectures do not require planetary flybys. To compare semi-cyclers and cyclers with the other architectures on a consistent basis, however, requires further optimization than what is available in the literature. For example, Mars-Earth semi-cycler trajectories have been computed previously, but not optimized across a range of TOF. A variety of cycler trajectories was also available, but the ΔV or TOF was not ideal for Mars mission scenarios. In the search for better cyclers, my research contributed to identifying two new families of cyclers for use in human mission to Mars.43-44 Finally, I have designed four versions of Earth-Mars semi-cycler trajectories to complete the assessment of Mars mission architectures.47 An example trajectory for each trajectory type is presented in Fig. 1–Fig. 4.

[pic]

Fig. 1 Outbound and inbound direct transfers.

[pic][pic][pic]

Fig. 2 Mars-Earth semi-cyclers.

[pic][pic]

Fig. 3 Earth-Mars semi-cyclers.

[pic]

Fig. 4 Outbound cycler trajectory with E1-E3 near 3:2 resonance and E3-E4 near 1.5 year transfer.

Should an accident on the way to Mars preclude the crew from landing (e.g. a propulsion system failure), a free-return trajectory68–75 would allow the crew to return to Earth without any major maneuvering (i.e. zero deterministic ΔV). These trajectories are constructed such that if there is no capture maneuver at arrival, a gravity assist from Mars will send the crew and vehicle back to Earth. I examine free-return trajectories for direct, semi-direct, and stop-over mission scenarios, though free-returns may be used (and are often incorporated) in semi-cycler and cycler scenarios. An example free-return trajectory is presented in Fig. 5. While the free-return abort is available, the nominal mission only uses the Earth-Mars portion of the trajectory. The crew would stay on Mars for about 550 days then take a short TOF inbound trajectory home (e.g. the Mars-Earth transit in Fig. 1).

[pic]

Fig. 5 Mars free-return trajectory with near 3:2 resonance from E1-E3.

I model the heliocentric trajectories as point-to-point conics with instantaneous V( rotations at planetary encounters. The minimum allowable flyby altitude at Earth and Mars is 300 km. Deep space maneuvers are also modeled as instantaneous changes in the heliocentric velocity. We do not allow maneuvers within the sphere of influence of a flyby planet because of the operational difficulty in achieving an accurate ΔV during a gravity assist. We assume that planetary departure and arrival maneuvers occur at 300 km above the planet’s surface, thus the ΔV for escape or capture is

[pic] (1)

where μ is the gravitational parameter of the planet and rp is the periapsis radius of the escape or capture hyperbola (in this case 300 km above the surface radius). While Eq. (1) is explicitly the ΔV to achieve a V( magnitude from a parabola, it is sufficient to optimize interplanetary transfers that begin on the surface or in a parking orbit. The difference between the true ΔV and Eq. (1) is found by subtracting the launch trajectory or parking orbit velocity at rp from the periapsis velocity of the parabola. Because this difference is independent of the interplanetary transfer (both the parking orbit and the parabola are planetocentric trajectories), it does not affect the outcome of the optimal trajectory.

The sequence of maneuvers included in the optimal ΔV calculation is summarized in Table 3 for each trajectory type. If the crew taxi, transfer vehicle (TV), or both vehicles performs a maneuver (when the maneuvers are required is provided in Table 3), the weighting on the corresponding ΔV is unity, and if no maneuver is performed then the weighting is zero. For this analysis, I assume the semi-cycler or cycler transfer vehicle is already in a parking orbit or on an interplanetary trajectory; thus, the initial transfer vehicle launch cost for these trajectories are ignored. While these relative weightings (one or zero) do not explicitly minimize IMLEO or cost, the resulting trajectories are representative of those that result from more detailed analyses (e.g. one that includes the vehicle masses). A key benefit of this weighting system is that I only rely on natural parameters (planetary orbits and masses) for computations, yet retain trajectory features (i.e. low V( and low ΔV) that are essential for effective integrated mission design.

Table 3 Required maneuvers for each trajectory type

|Trajectory |Earth |Mars |Mars |Earth |DSM |

| |departure ΔV |arrival ΔV |departure ΔV |arrival ΔV | |

|Direct |taxi & TV |TVa or neitherb |taxi & TV |TVa or neitherb |neither |

|Free-return |taxi & TV |TVa or neitherb |N/A |N/A |neither |

|M-E semi-cyclerc |taxi |TVa or neitherb |taxi & TV |neither |TV |

|E-M semi-cyclerc |taxi & TV |neither |taxi |TVa or neitherb |TV |

|Cyclerc |taxi |neither |taxi |neither |TV |

aPowered capture.

bAero-assisted capture.

cThe one-time transfer vehicle launch ΔV is ignored.

I use a sequential quadratic programming algorithm76,77 to compute minimum-ΔV trajectories with bounded TOF. (By bounded I mean the TOF may be less than or equal to the constrained value.) Similar methods have been used in the optimization of the Galileo trajectory to and at Jupiter and the Cassini trajectory to Saturn.78, 79 I have developed my own software that provides a user-defined objective function and constraints (with gradients) to a commercial (MATLAB) optimizer. Earth-Mars trajectories are optimized so that the total ΔV over the entire 15-year cycle is minimized (as opposed to, say, minimizing the maximum ΔV during the cycle). Though the arrival V( for aerocapture is often limited (e.g. below 9 km/s at Earth65 and below 7 km/s at Mars66), I did not constrain the V(; in this way the lowest possible ΔV trajectories are analyzed.

An initial guess for the timing and placement of deep-space maneuvers (DSMs) is obtained via Lawden’s primer vector analysis.80, 81 For example, in Fig. 6 a DSM would be placed where the primer magnitude is largest (in this case around 7.5 years time of flight). An augmented trajectory (with the new DSM) is then optimized for minimum ΔV . Additional DSMs are added until a locally optimal trajectory is found. Fig. 7 demonstrates an optimal trajectory, where the primer vector P satisfies the five conditions for optimality:

1. P and [pic] are continuous.

2. P is aligned with ΔV at impulse times.

3. [pic] at impulse times.

4. [pic] on coasting arcs separating impulses.

5. [pic]at impulses.

[pic]

Fig. 6 Primer vector magnitude along sub-optimal cycler trajectory.

[pic]

Fig. 7 Primer vector magnitude along optimal cycler trajectory.

Using the SQP algorithm in concert with primer vector analysis, I computed optimal-ΔV direct, free-return, Mars-Earth semi-cycler, Earth-Mars semi-cycler, and cycler trajectories with the TOF constrained to below 120 to 270 days for both powered and aero-assisted planetary capture.82 A synopsis of the results is presented in Fig. 8 for the average ΔV over seven consecutive missions and in Fig. 9 for the maximum ΔV over the seven-mission timescale.

[pic][pic]

Determining the maximum payload for low-thrust transfers83–95 requires optimization of the propulsion system along with the trajectory. For example, a increase in thrust lowers the ΔV (and propellant mass fraction) for the vehicle, but the hardware mass must increase to provide additional power to the thrusters. Consequently, there is a trade between low propellant mass with high power levels and low hardware mass at low power levels. The balance in this trade is usually determined by the specific mass (α) and efficiency (η) of the propulsion system. The specific mass determines the ratio between the hardware (power source, thrusters, etc.) and the amount of power it produces

[pic] (2)

and the efficiency is the ratio of jet power to hardware power

[pic] (3)

The jet power may be determined from

[pic] (4)

Thus, the hardware mass is

[pic] (5)

and, for constant α and η, increases proportionally with the thrust and specific impulse. The propellant tanks also contribute a significant portion of the spacecraft mass, and the tank mass may be estimated via the tankage factor

[pic] (6)

The final-to-initial mass fraction may be computed for low-thrust missions via the rocket equation96

[pic] (7)

The mass ratio may be written explicitly as:

[pic] (8)

where we assume that the thrust and specific impulse are constant and that the thrust may be computed by

[pic] (9)

The rocket equation is particularly useful in determining the ratio of initial mass (payload, hardware, tankage, and propellant) to payload mass

[pic] (10)

The spacecraft thrust is thus

[pic] (11)

In order to minimize the initial mass [Eq. (10)] or thrust [Eq. (11)], an accurate means of determining the minimum ΔV for a given trajectory TOF and vehicle a0 and Isp is required. Zola85 describes an approximate analytic method to calculate ΔV as a function of a0 and Isp, provided the ΔV-optimal burn time tb for a trajectory with the same TOF is known. In contrast to the maximum-acceleration trajectories provided in Fig. 8 and Fig. 9, I present the ΔV for a direct transfer from Earth to Mars (and vice-versa) using the minimum possible acceleration during launch years 2009–2022 in Fig. 10. In the specific case of Fig. 10, I calculated trajectories that match the heliocentric orbit of Mars from the orbit of Earth and set the mass flow rate to zero (corresponding to infinite Isp). Because the acceleration is minimized, the burn time for these trajectories is equal to the TOF (i.e. the thruster is always on). Using this data, the heuristic method of Ref. 85 produces values for optimum payload mass fractions to within a few percent. Alternatively, numerical optimization techniques provide higher fidelity results at the expense of longer computation time.

I also examined minimum-acceleration transfers for aerocapture missions (for arrival [pic]) but found that the V( at arrival was impractical (i.e. it is often at least double the impulsive transfer V(). Instead, I optimized low-thrust trajectories with infinite Isp (i.e. assuming constant mass) and set the acceleration to the levels found in Fig. 10 for a given launch year and TOF combination. The resulting ΔV are found in Fig. 11. Because the same initial acceleration was used to calculate the powered and aero-assisted capture trajectories, the burn time is given by tb = (ΔVA/ΔVP)TOF where ΔVP is the powered arrival ΔV (from Fig. 10) and ΔVA is the aero-arrival ΔV (from Fig. 11). Generally, increasing the acceleration decreases the ΔV and burn time, and in the limit of infinite thrust and zero burn time, the ΔV approaches the sum of the departure and arrival V( for powered capture missions and becomes the departure V( for aero-assisted capture.

[pic] [pic]

Fig. 10 ΔV for minimum-thrust transfers with powered capture.

[pic][pic]

Fig. 11 ΔV for low-thrust transfers with aerocapture or direct entry.

Parking Orbit Reorientation

Because parking orbits are an essential ingredient to semi-direct, stop-over, and semi-cycler mission scenarios, an effective means of capturing into then departing from an orbit is required.97–103 In many cases the parking orbit is not conveniently oriented with respect to the interplanetary leg (e.g., returning to Earth from a parking orbit at Mars). One solution to this problem is to reorient the spacecraft’s orbit about the planet before departure. This method includes a maneuver at apoapsis that rotates the parking orbit about the line of apsides to achieve the proper orientation at departure, thus coupling the effects of parking-orbit orientation with the interplanetary trajectories. This method is thus termed the “apo-twist” maneuver because the parking orbit is “twisted” about the line of apsides via a ΔV at apoapsis.101, 103 I also account for the natural precession of the parking orbit (due to J2, lunar, and solar perturbations) during the stay time.

Fig. 12 Orientation of required[pic]and parking orbit.

Fig. 12 illustrates the problem of entering into and departing from a parking orbit by presenting an ideal situation (Case 1) and a non-ideal situation (Case 2). Case 1 presents no difficulties because the planet provides sufficient bending to reach [pic] with a tangential burn at periapsis. Here, the approach hyperbola, parking orbit, and departure hyperbola are coplanar and have the same periapsis. To calculate the ΔV (assuming conic trajectories) only the magnitudes of [pic] are needed; the cost of entering the parking orbit is found explicitly via Eq. (12),

[pic] (12)

where μ is the gravitational parameter of the planet, rp is the radius of periapsis, and a is the semi-major axis of the parking orbit. The cost of departing is found by replacing [pic] with [pic].

To achieve [pic] in Case 2 seems difficult, but is possible with the apo-twist maneuver. In this case, [pic], [pic], and the parking orbit do not necessarily lie in the same plane. Even in the case where [pic] and [pic] are in the plane of the page, it will be necessary to target [pic] so that the arrival parking orbit is not in the plane of the page. The selection of the arrival parking orbit plane, combined with a maneuver at apoapsis (to rotate the departure parking orbit along the line of apsides) are the key ingredients of the apo-twist method. With this technique, the approach, parking orbit, and departure trajectories share a common periapsis, but the total ΔV will be greater than that provided by Eq. (12). The apo-twist is illustrated in Fig. 13, where [pic], the arrival orbit, and [pic] (without apo-twist) are in the plane of the page. The desired [pic] (which is out of the plane of the page) is achieved by rotating the parking orbit via the apo-twist maneuver.

Fig. 13 Rotation of parking orbit about line of apsides by twist angle [pic].

To assess the effectiveness of the apo-twist maneuver (which requires three burns), I compare it to the optimal two-burn scheme. The first burn is the orbit insertion maneuver, which is performed tangentially to the velocity at periapsis. Thus, the insertion maneuver may be replaced by atmospheric braking (aerocapture) if desired. The second burn is the departure maneuver. Generally, this is a three-dimensional maneuver (i.e. not tangential at periapsis) because the parking orbit and [pic] are not perfectly aligned (Case 2 of Fig. 12).

The minimum ΔV and corresponding inclination, using both the apo-twist and two-burn methods, are provided in Table 4 and Table 5. The optimal time to perform the apo-twist maneuver for each arrival date is found under the tΔV column in these tables. We note that an apo-twist maneuver is generally available at any point in the stay time, but the lowest ΔV maneuvers often occur either at the beginning or end of the stay time. The [pic] and [pic] columns give the cost of arriving and departing the orbit with a tangential periapsis burn using Eq. (12). We note that in the case of aerocapture the propulsive [pic]. The [pic] columns under “Apo-twist” and “Two-burn” provide the additional cost of achieving the correct direction at departure for these methods. Thus the total cost is

[pic] (13)

Since the reorientation cost can be a significant fraction of the total ΔV, it cannot be neglected in mission design. Moreover, an economical way of reducing this reorientation ΔV will lower the propellant cost of the mission. I find that the apo-twist maneuver requires the least ΔV when compared to other methods.

Table 4 Stop-over cycler,38 Mars

| | |Ideala | |Apo-twist | |Two-burn |

|Arrival Date |Stay time |ΔVA (km/s) |Δ| |tΔ|iA (deg) |

| |(days) | |V| |V | |

| | | |D| |(d| |

| | | |(| |ay| |

| | | |k| |s)| |

| | | |m| | | |

| | | |/| | | |

| | | |s| | | |

| | | |)| | | |

|Arrival Date |Stay time (days) |ΔVA (km/s) |ΔVD (km/s) |

|Hydrogen/Oxygen (H2/O2) |450 |0.15 |0.10 |

|Methane/Oxygen (CH4/O2) |380 |0.12 |0.08 |

|Nuclear (NTR) |900 |0.50 |0.15 |

|Cargo NEP |8,000 |10–30 kg/kW |0.15 |

|Transfer Vehicle NEP |3,000–5,000 |10 kg/kW |0.15 |

The initial-to-payload mass fraction is determined for a given maneuver for impulsive ΔV by

[pic] (14)

and for low-thrust transfers via

[pic] (15)

The sequence of stages may be combined and stacked to produce the IMLEO.

The payload mass depends on the assumed crew number, cargo mass, and vehicle masses. I plan to normalize the IMLEO on a per person basis (i.e. IMLEO is given in mt/person). Nominally the taxi and transfer vehicle masses are 1.25 mt/person and 5 mt/person, respectively. The cargo mass varies from 0–10 mt/person, and I also vary the taxi mass from 1–3 mt/person and the transfer vehicle from 3–10 mt/person.

Given these parameters (along with a few other assumptions) I can provide IMLEO tables of propulsion technology versus architecture for a given cargo mass, taxi, transfer vehicle, and TOF. I can also produce plots of IMLEO versus TOF or transfer vehicle mass. Finally, I can produce charts of the IMLEO-optimal architecture as a function of transfer vehicle mass and TOF.

Once the IMLEO data is collected, the technologies that provide the greatest IMLEO reduction for a given development investment should emerge (e.g. Martian water may provide the best IMLEO, but NTR reduces IMLEO with less development). I also plan to provide recommendation for how to evolve from early development phases to colonization phases of exploration. Also, as transfer vehicle mass increases to allow more safety and comfort to the crew then the optimal architecture will likely evolve from semi-direct to cycling concepts, but the best path or optimal architecture is still unknown.

References

1. Von Braun, W., The Mars Project, University of Illinois Press, Urbana, IL, 1953.

2. Stuhlinger, E., “Electrical Propulsion System for Space Ships with Nuclear Power Source,” Journal of the Astronautical Sciences, Part I, Vol. 2, winter 1955, pp.149–152; Part II, Vol. 3, spring 1956, pp.11–14; Part III, Vol. 3, summer 1956, p.33.

3. Ehricke, K. A., Whitlock, C. M., Chapman, R. L., and Purdy, C. H., “Calculations on a Manned Nuclear Propelled Space Vehicle,” ARS Report 532–57, 1957.

4. Himmel, S. C., Dugan, J. F., Luidens, R. W., and Weber, R. J., “A Study of Manned Nuclear-Rocket Missions to Mars,” Aerospace Engineering, Vol. 20, July 1961, pp. 18, 19, 51–58.

5. Gillespie, R. W., Ragsac, R. V., Ross, S. E., “Prospects for Early Manned Interplanetary Flights,” Astronautics and Aerospace Engineering, Vol. 1, August 1963, pp. 16–21.

6. Dixon, F. P., “The EMPIRE Dual Planet Flyby Mission,” AIAA/NASA Conference on Engineering Problems of Manned Interplanetary Travel, Palo Alto, CA, September 30–October 1, 1963, pp. 3–18.

7. Titus, R. R., “FLEM–Flyby-Landing Excursion Mode,” AIAA Paper 66-36, Aerospace Sciences Meeting, New York, NY, July 26–29, 1965.

8. Von Braun, W., “The Next 20 Years of Interplanetary Exploration,” Astronautics and Aeronautics, Vol. 3, November 1965, pp. 24–34.

9. King, J. C., Shelton, R. D., Stuhlinger, E., and Woodcock, G. R., “Study of a Nerva-Electric Manned Mars Vehicle,” AIAA/AAS Stepping Stones to Mars Meeting, Baltimore, MD, March 28–30, 1966, pp. 288–301.

10. Bell, M. W. J., “An Evolutionary Program for Manned Interplanetary Exploration,” Journal of Spacecraft and Rockets, Vol. 4, No. 5, 1967, pp. 625–630.

11. Farquhar, R. W., “Future Missions for Libration-Point Satellites,” Astronautics and Aeronautics, Vol. 7, May 1969, pp. 52–56.

12. Mueller, G. E., “An Integrated Space Program for the Next Generation,” Astronautics and Aeronautics, Vol. 8, January 1970, pp. 30–51.

13. Cohen, A., The 90 Day Study on the Human Exploration of the Moon and Mars, U.S. Government Printing Office, Washington, DC, 1989.

14. Zubrin, R., Baker, D., and Gwynne, O., “Mars Direct: A Simple, Robust, and Cost-Effective Architecture for the Space Exploration Initiative,” AIAA Paper 91–0326, Aerospace Sciences Meeting, Reno, NV, January 1991.

15. Ordway, F., Sharpe, M., and Wakeford, R., “EMPIRE: Early Manned Planetary-Interplanetary Roundtrip Expeditions Part I: Aeronutronic and General Dynamics Studies,” Journal of the British Interplanetary Society, Vol. 46, 1993, pp. 179–190.

16. Ordway, F., Sharpe, M., and Wakeford, R., “EMPIRE: Early Manned Planetary-Interplanetary Roundtrip Expeditions Part II: Lockheed Missiles and Space Studies,” Journal of the British Interplanetary Society, Vol. 47, 1994, pp. 181–190.

17. Donahue, B. B., “Mars Ascent-Stage Design Utilizing Nuclear Propulsion,” Journal of Spacecraft and Rockets, Vol. 32, No. 3, May-June 1995, pp. 552–558.

18. Zubrin, R. and Wagner, R., The Case for Mars, Simon and Schuster, Inc., New York, NY, 1996.

19. Hoffman, S. and Kaplan, D., eds., “Human Exploration of Mars: The Reference Mission of the NASA Mars Exploration Study Team,” NASA SP 6107, 1997.

20. Drake, B. G., ed., “Reference Mission Version 3.0 Addendum to the Human Exploration of Mars: The Reference Mission of the NASA Mars Exploration Study Team,” Exploration Office Document EX 13-98-036, June 1998.

21. Burgess, E., Rocket Propulsion: with an Introduction to the Idea of Interplanetary Travel, Chapman and Hall, London, 1952, pp. 172–179.

22. Irving, J. H. and Blum, E. K., “Comparative Performance of Ballistic and Low-Thrust Vehicle for Flight to Mars,” Vistas in Astronautics, Vol. II, Pergamon Press, 1959, pp.191–218.

23. Chovit, A. R., Callies, G. M., Cannon, G. W., and Simmons, L. D., “Manned Interplanetary Mission Modes for Nuclear Propulsion Systems,” AIAA Paper 68-588, AIAA 4th Propulsion Joint Specialist Conference, Cleveland, OH, June 10–14, 1968.

24. Hoffman, S. J., Friedlander, A. L., and Nock, K. T., “Transportation Mode Performance Comparison for a Sustained Manned Mars Base,” AIAA Paper 86-2016, AIAA/AAS Astrodynamics Conference, Williamsburg, VA, August 18–20, 1986.

25. Adams, A., Priest, C., Sumrall, P., and Woodcock, G., “Overview of Mars Transportation Options and Issues,” AIAA Paper 90-3795, AIAA Space Programs and Technologies Conference, Huntsville, AL, September 25–28, 1990.

26. Cothran, B., Tillotson, B., and Donahue, B., “Advanced Propulsion Options for Human Exploration of Mars,” AIAA Paper 90-3822, AIAA Space Programs and Technologies Conference, Huntsville, AL, September 25–28, 1990.

27. Braun, R. D. and Blersch, D. J., “Propulsive Options for a Manned Mars Transportation System,” Journal of Spacecraft and Rockets, Vol. 28, No. 1, 1993, pp. 85–92.

28. Walberg, G., “How Shall We Go to Mars? A Review of Mission Scenarios,” Journal of Spacecraft and Rockets, Vol. 30, No. 2, 1993, pp. 129–139.

29. Niehoff, J. C. and Hoffman, S. J., “Pathways to Mars: An Overview of Flight Profiles and Staging Options for Mars Missions,” AAS Paper 95-478, Science and Technology Series of the American Astronautical Society, Vol. 86, San Diego, CA, Univelt, Inc., 1996, pp. 99–125.

30. Donahue, B. B. and Cupples, M. L., “Comparative Analysis of Current NASA Human Mars Mission Architectures,” Journal of Spacecraft and Rockets, Vol. 38, No. 5, 2001, pp.745–751.

31. Landau, D. F. and Longuski, J. M., “Comparative Assessment of Human Missions to Mars,” AAS Paper 03-513, AAS/AIAA Astrodynamics Specialist Conference, Big Sky, MT, August 4–7, 2003.

32. Griffin, B., Thomas, B., Vaughan, D., Drake, B., Johnson, L., and Woodcock, G., “A Comparison of Transportation Systems for Human Missions to Mars,” AIAA Paper 2004-3834, 40th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Fort Lauderdale, FL, July 11–14, 2004.

33. Bounova, G. A., Ahn, J., Hofstetter, W., Wooster, P., Hassan, R., and de Weck, O. L., “Selection and Technology Evaluation of Moon/Mars Transportation Architectures,” AIAA paper 2005-6790, Space 2005, Long Beach, CA, August 30–September 1, 2005.

34. Ramohalli, K. N. R., and Marcozzi, M., “Cost reductions through ISRU - Recent developments,” IAA Paper 97-1505, 48th International Astronautical Congress, Turin, Italy, October 6–10, 1997

35. Larson, W. J., and Pranke, L. K., Human Spaceflight: Mission Analysis and Design, McGraw-Hill, New York, 1999.

36. Hunt, C. D., and vanPelt, M. O., “Comparing NASA and ESA Cost Estimating Methods for Human Missions to Mars,” 26th International Society of Parametric Analysts Conference, Frascati, Italy, May 10–12, 2004.

37. Hall, J. L., Noca, M. A., and Bailey, R. W., “Cost-Benefit Analysis of the Aerocapture Mission Set,” Journal of Spacecraft and Rockets, Vol 42, No. 2, 2005, pp.309–320.

38. Niehoff, J., Friedlander, A., and McAdams, J., “Earth-Mars Transport Cycler Concepts,” International Astronautical Federation, IAF Paper 91-438, International Astronautical Congress, Montreal, Canada, October 5–11, 1991.

39. Penzo, P. and Nock, K., “Earth-Mars Transportation Using Stop-Over Cyclers,” AIAA Paper 2002-4424, AIAA/AAS Astrodynamics Specialist Conference, Monterey, CA, August 2–5 2002.

40. Hollister, W. M, “Castles in Space,” Astronautica Acta, Vol. 14, No. 2, 1969, pp. 311-316.

41. Rall, C. S. and Hollister, W. M., “Free-Fall Periodic Orbits Connecting Earth and Mars,” AIAA Paper 71-92, 1971.

42. Friedlander, A. L., J. C. Niehoff, D. V. Byrnes, and J. M. Longuski, “Circulating Transportation Orbits Between Earth and Mars,” AIAA Paper 1986-2009, AIAA/AAS Astrodynamics Conference, Williamsburg, VA, August 18–20, 1986.

43. Byrnes, D. V., Longuski, J. M., and Aldrin, B., “Cycler Orbit Between Earth and Mars,” Journal of Spacecraft and Rockets, Vol. 30, No. 3, May-June 1993, pp. 334–336.

44. Byrnes, D. V., McConaghy, T. T, and Longuski, J. M., “Analysis of Various Two Synodic Period Earth-Mars Cycler Trajectories,” AIAA Paper 2002-4423, AIAA/AAS Astrodynamics Specialist Conference, Monterey, CA, August 5–8, 2002.

45. McConaghy, T. T., Yam, C. H., Landau, D. F., and Longuski, J. M., “Two-Synodic-Period Earth-Mars Cyclers with Intermediate Earth Encounter,” AAS Paper 03-509, AAS/AIAA Astrodynamics Specialist Conference, Big Sky, MT, August 4–7, 2003. To appear in the Journal of Spacecraft and Rockets.

46. Russell, R. P., and Ocampo, C. A., “Global Search for Idealized Free-Return Earth-Mars Cyclers,” Journal of Guidance, Control, and Dynamics, Vol. 28, No. 5, March-April 2005, pp. 194–208.

47. Chen, K. J., McConaghy, T. T., Landau, D. F., and Longuski, J. M., and Aldrin, B., “Powered Earth-Mars Cycler with Three Synodic-Period Repeat Time,” Journal of Spacecraft and Rockets, Vol. 42, No. 5, September–October 2005, pp. 921–927.

48. Bishop, R. H., Byrnes, D. V., Newman, D. J., Carr, C. E., and Aldrin, B., “Earth-Mars Transportation Opportunities: Promising Options for Interplanetary Transportation,” American Astronautical Society, AAS Paper 00-255, The Richard H. Battin Astrodynamics Conference, College Station, TX, March 20-21, 2000.

49. Aldrin, B., Byrnes, D., Jones, R., and Davis, H., “Evolutionary Space Transportation Plan for Mars Cycling Concepts,” AIAA Paper 2001-4677, Albuquerque, NM, August 2001.

50. Landau, D. F., and Longuski, J. M., “Mars Exploration via Earth-Mars Semi-Cyclers,” AAS Paper 05-269, Lake Tahoe, CA, August 7–11, 2005.

51. Moorhouse, D. J., “Detailed Definitions and Guidance for Application of Technology Readiness Levels,” Journal of Aircraft, Vol. 39, No. 1, 2002, pp. 190–192.

52. Sohn, R. L., “Venus Swingby Mode for Manned Mars Missions,” Journal of Spacecraft and Rockets, Vol. 1, No. 5, 1964, pp. 565–567.

53. Deerwester, J. M. and Dhaem, S. M., “Systematic Comparison of Venus Swingby Mode with Standard Mode of Mars Round Trips,” Journal of Spacecraft and Rockets Vol. 4, July 1967, pp. 904–911.

54. Gillespie, R. W. and Ross, S., “Venus-Swingby Mode and Its Role in the Manned Exploration of Mars,” Journal of Spacecraft and Rockets Vol. 4, February 1967, pp. 170–175.

55. Casalino, L., Colasurdo, G., Pastrone, D., “Optimization Procedure for Preliminary Design of Opposition-Class Mars Missions,” Journal of Guidance, Control, and Dynamics, Vol. 21, No. 1, 1998, pp. 134–140.

56. Breakwell, J. V., Gillespie, R. W., and Ross, S. E., “Researches in Interplanetary Flight,” ARS Journal, Vol. 31, No. 2, 1961, pp.201–207.

57. Knip, G. and Zola, C. L., “Three-Dimensional Trajectory Analysis for Round-Trip Mission to Mars,” NASA TN-D-1316, 1962.

58. Ross, S., “Planetary Flight Handbook,” NASA SP-35, 1963.

59. Lee, V. A. and Wilson, S. W., “A Survey of Ballistic Mars-Mission Profiles,” Journal of Spacecraft and Rockets, Vol. 4, No. 2, 1967, pp. 129–142.

60. Young, A. C., Mulqueen, J. A., and Skinner, J. E., “Mars Exploration, Venus Swingby and Conjunction Class Mission Modes, Time Period 2000-2045,” NASA TM-86477, 1984.

61. Gravier, J. P., Marchal, C., Culp, R. D., “Optimal Trajectories Between Earth and Mars in Their True Planetary Orbits,” Journal of Optimization Theory and Applications, Vol. 9, February 1972, pp. 120–136.

62. Hoffman, S. J., McAdams, J. V., and Niehoff, J. C., “Round Trip Trajectories for Human Exploration of Mars,” AAS Paper 89-201, 1989.

63. Soldner, J. K., “Round Trip Mars Trajectories: New Variations on Classic Mission Profiles,” AIAA Paper 90-2932, August 1990.

64. George, L. E. and Kos, L. D., “Interplanetary Mission Design Handbook: Earth-to Mars Mission Opportunities and Mars-to-Earth Return Opportunities 2009-2024,” NASA TM-1998-208533, 1998.

65. Munk, M. M., “Departure Energies, Trip Times, and Entry Speeds for Human Mars Missions,” American Astronautical Society, AAS Paper 99-103, 1999.

66. Miele, A. and Wang, T., “Optimal Transfers from an Earth Orbit to a Mars Orbit,” Acta Astronautica, Vol. 45, No. 3, 1999, pp. 119–133.

67. Braun, R. D., Powell, R. W., and Hartung, L. C., “Effect of Interplanetary Options on a Manned Mars Aerobrake Configuration,” NASA TP-3019, August 1990.

68. Striepe, S. A., Braun, R. D., Powell, R. W., and Fowler, W. T., “Influence of Interplanetary Trajectory Selection on Earth Atmospheric Velocity of Mars Missions,” Journal of Spacecraft and Rockets, Vol. 30, No. 4, 1993, pp. 420–425.

69. Striepe, S. A., Braun, R. D., Powell, R. W., and Fowler, W. T., “Influence of Interplanetary Trajectory Selection on Mars Atmospheric Velocity,” Journal of Spacecraft and Rockets, Vol. 30, No. 4, 1993, pp. 426–430.

70. Lyne, J. E., Wercinski, P., Walberg, G., and Jits, R., “Mars Aerocapture Studies for the Design Reference Mission,” AAS Paper 98-110, AAS/AIAA Space Flight Mechanics Meeting, Monterey, CA, February 9–11, 1998.

71. Crocco, G. A., “One Year Exploration Trip Earth-Mars-Venus-Earth,” Proceedings of the 10th International Astronautical Federation, Rome, 1956.

72. Battin, R. H, “The Determination of Round-Trip Planetary Reconnaissance Trajectories,” Journal of Aerospace Sciences, Vol. 26, No. 9, 1959, pp. 545–567.

73. Ruppe, H. O., “Interplanetary Flight,” Handbook of Astronautical Engineering, H. H. Koelle, ed., McGraw-Hill Book Company, New York, 9.32–9.44, 1961.

74. Hénon, M., “Interplanetary Orbits which Encounter the Earth Twice,” Bulletin Astronomique, Vol. 3, No. 3, 1968, pp.377–393.

75. Wolf, A., A., “Free Return Trajectories for Mars Missions,” AAS Paper 91-123, AAS/AIAA Spaceflight Mechanics Meeting, San Diego, CA, February 11–13, 1991.

76. Patel, M. R., Longuski, J. M., and Sims, J. A., “Mars Free Return Trajectories,” Journal of Spacecraft and Rockets Vol. 35, No. 3, 1998, pp. 350–354.

77. Miele, A., Wang, T., and Mancuso, S., “Optimal Free-Return Trajectories for Moon Missions and Mars Missions,” Journal of the Astronautical Sciences, Vol. 48, Nos. 2-3, 2000, pp.183–206.

78. Okutsu, M. and Longuski, J. M., “Mars Free-Returns via Gravity Assist from Venus,” Journal of Spacecraft and Rockets Vol. 39, No. 1, 2002, pp. 31–36.

79. Optimization Toolbox User's Guide, The MathWorks, Inc., 2004.

80. Gill, P. E., Murray, W., and Saunders, M. A., “SNOPT: An SQP Algorithm for Large-Scale Constrained Optimization,” SIAM Journal on Optimization, Vol. 12, 2002, pp. 979–1006.

81. D’Amario, L., A., Byrnes, D. V., Sackett, L. L., and Stanford, R. H., “Optimization of Multiple Flyby Trajectories,” AAS Paper 79-162, AAS/AIAA Astrodynamics Specialist Conference, Provincetown, MA, June 25–27, 1979.

82. Sauer, C. G., “Optimization of Interplanetary Trajectories with Unpowered Planetary Swingbys,” AAS Paper 87-424, AAS/AIAA Astrodynamics Specialist Conference, Kalispell, MT, August 10–13 1987.

83. Lawden, D. F., Optimal Trajectories for Space Navigation, Butterworths, London 1963, Ch. 3, pp. 55–69, 63.

84. Jezewski, D.J., “Primer Vector Theory and Applications,” NASA TR-R-454, November 1975.

85. Landau, D. F. and Longuski, J. M., “A Reassessment of Trajectory Options for Human Missions to Mars,” AIAA Paper 2004-5095, AIAA/AAS Astrodynamics Specialist Conference, Providence, RI, August 16–19, 2004.

86. Moekel, W. E., “Fast Interplanetary Missions with Low-Thrust Propulsion Systems,” NASA TR-R-79, 1961.

87. Melbourne, W. G. and Sauer, C. G., “Optimum Interplanetary Rendezvous with Power-Limited Vehicles,” AIAA Journal, Vol. 1, No. 1, 1963, pp. 54–60.

88. Zola, C. L., “A Method for Approximating Propellant Requirements of Low-Thrust Trajectories,” NASA TN-D-3400, 1966.

89. Ragsac, R. V., “Study of Electric Propulsion for Manned Mars Missions,” Journal of Spacecraft and Rockets, Vol. 4, April 1967, pp. 462–468.

90. Kawaguchi, J., Takiura, K., and Matsuo, H., “On the Optimization and Application of Electric Propulsion to Mars and Sample and Return Mission,” AAS Paper 94-183, AAS/AIAA Spaceflight Mechanics Meeting, Cocoa Beach, FL, February 1994.

91. Chang-Diaz, F. R., Hsu, M. M., Braden, E., Johnson, I., Yang, T. F., “Rapid Mars Transits with Exhaust Modulated Plasma Propulsion,” NASA TP-3539, 1995.

92. Tang, S. and Conway, B. A., “Optimization of Low-Thrust Interplanetary Trajectories Using Collocation and Nonlinear Programming,” Journal of Guidance, Control, and Dynamics, Vol. 18, No. 3, 1995, pp. 599–604.

93. Frisbee, R. H., and Hoffman, N. J., “Electric Propulsion Options for Mars Cargo Missions,” AIAA Paper 96-3173, 32nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Lake Buena Vista, FL, July 1–3, 1996.

94. Gefert, L. P., Hack, K. J., and Kerslake, T. W., “Options for the Human Exploration of Mars using Solar Electric Propulsion,” STAIF, Proceedings of the Conferences on Applications of Thermophysics in Microgravity and on Next Generation Launch Systems, and 16th Symposium on Space Nuclear Power and Propulsion, Albuquerque, NM, January 31-February 4, 1999, pp. 1275–1280.

95. Williams, S. N. and Coverstone-Carrol, V., “Mars Missions Using Solar-Electric Propulsion,” Journal of Spacecraft and Rockets, Vol. 37, No. 1, 2000, pp.71–77.

96. McConaghy, T. T., Debban, T. J., Petropoulus, A. E., and Longuski, J. M., “An Approach to Design and Optimization of Low-thrust Trajectories with Gravity Assists,” AAS Paper 01-468, AAS/AIAA Astrodynamics Specialist Conference, Quebec City, QC, Canada, July-August 2001.

97. Whiffen, G. J. and Sims, J. A., “Application of the SDC Optimal Control Algorithm to Low-Thrust Escape and Capture Trajectory Optimization,” AAS Paper 02-208, AAS/AIAA Space Flight Mechanics Conference, San Antonio, TX, January 27–30, 2002.

98. Sankaran, K., Cassady L., Kodys, A. D., and Choueiri, E. Y., “A Survey of Propulsion Options for Cargo and Piloted Missions to Mars,” The Annals of the New York Academy of Science, Vol. 1017, May 2004, pp. 450–467.

99. Tsiolkovsky, K. E., “Exploration of the Universe with Reaction Machines,” The Science Review, #5, St. Petersburg, Russia, 1903.

100. Desai, P., Braun, R., and Powell, R., “Aspects of Parking Orbit Selection in a Manned Mars Mission,” NASA TP-3256, Washington, D.C., December 1992.

101. Desai, P., and Bugulia, J., “Arrival and Departure Impulsive ΔV Determination for Precessing Mars Parking Orbits,” Journal of the Astronautical Sciences, Vol. 41, No.1, January-March 1993, pp. 1–18.

102. Thibodeau III, J. R., “Use of Planetary Oblateness for Parking-Orbit Alignment,” NASA TN D-4657, Washington, D.C., July 1968.

103. Desai, P., and Bugulia, J., “Determining Mars Parking Orbits that Ensure Tangential Periapsis Burns at Arrival and Departure,” Journal of Spacecraft and Rockets, Vol. 30, No. 4, July-August 1993, pp. 414–419.

104. Luidens, R. W., and Miller, B. A., “Efficient Planetary Parking Orbits with Examples for Mars,” NASA TN D-3220, Washington, D.C., January 1966.

105. Cupples, M. L., and Nordwall, J. A., “Optimal Parking Orbits For Manned Mars Missions,” AAS Paper 93-149, AAS/AIAA Spaceflight Mechanics Meeting, Pasadena, CA, February 22–24, 1993.

106. Landau, D. F., Longuski, J. M., and Penzo, P. A., “Method for Parking-Orbit Reorientation for Human Missions to Mars,” Journal of Spacecraft and Rockets, Vol. 42, No. 3, May-June 2005, pp. 517–522.

107. Kluever, C. A., “Feedback Control for Spacecraft Rendezvous and Docking,” Journal of Guidance, Control, and Dynamics, Vol. 22, No. 4, July–August 1999, pp.609–611.

108. Penzo, P. A. and Nock, K. T., “Hyperbolic Rendezvous for Earth-Mars Cycler Missions,” Paper AAS 02-162, AAS/AIAA Space Flight Mechanics Meeting, San Antonio, TX, January 27–30, 2002, pp. 763–772.

109. Wang, P. K. C., “Non-linear guidance laws for automatic orbital rendezvous,” International Journal of Control, Vol. 42, No. 3, 1985, pp. 651–670.

-----------------------

c) E2-E4 1.5 year transfer

b) E2-E3 near 2:3 resonance

a) E2-E3 near 2:1 Earth:spacecraft resonance

b) E1-E3 near 3:2 resonance, E3-E4 1:1 resonance, E4-E6 near 3:2 resonace

a) M2-M3 near 3:4 resonance

Fig. 9 Maximum DðV over 15-year cycle.

Fig. 8 Average DðV over 15-year cycle.

b) Mars to Earth

a) Earth to MaIC 9 Maximum ΔV over 15-year cycle.

Fig. 8 Average ΔV over 15-year cycle.

b) Mars to Earth

a) Earth to Mars

b) Mars to Earth

a) Earth to Mars

[pic]

[pic]

[pic]

[pic]

Case 1

Case 2

Twist angle, φ

[pic]

[pic]

Arrival orbit

(in plane of page)

Departure orbit

(out of plane of page)

φ

φ

[pic]

A (hyperbolic half-angle)

taxi

transfer

vehicle

from Mars

Earth

flyby

................
................

In order to avoid copyright disputes, this page is only a partial summary.

Google Online Preview   Download