Next Article in Journal
Earthwork Volume Calculation, 3D Model Generation, and Comparative Evaluation Using Vertical and High-Oblique Images Acquired by Unmanned Aerial Vehicles
Previous Article in Journal
Physics-Informed MTA-UNet: Prediction of Thermal Stress and Thermal Deformation of Satellites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Manned Mars Mission Analysis Using Mission Architecture Matrix Method

System Simulation Technology Laboratory (SST), School of Astronautics, Beihang University, Beijing 100191, China
*
Author to whom correspondence should be addressed.
Aerospace 2022, 9(10), 604; https://doi.org/10.3390/aerospace9100604
Submission received: 30 August 2022 / Revised: 19 September 2022 / Accepted: 20 September 2022 / Published: 14 October 2022
(This article belongs to the Section Astronautics & Space Science)

Abstract

:
With the development of deep space exploration technology, manned missions to Mars are expected to be realized in the near future. However, the journey to Mars requires more supplies and fuel than near-Earth missions, including moon landings. Using the traditional Apollo-style mission method will result in the spacecraft reaching the LEO mass of 2000 tons, which is not easy to achieve. The use of the modular design method and multiple launches will significantly reduce the mass of a single launch. In addition, the use of nuclear power engines (Nuclear Thermal Propulsion, NTP, and Nuclear Electric Propulsion, NEP) can greatly improve propulsion efficiency, reducing the mass of the propellant. This paper uses the Mission Architecture Matrix (MAM) method to concisely and precisely analyze a series of mission architectures in the case of impulse maneuver transfer and low-thrust transfer. The results show that when only chemical propulsion (specific impulse is 440 s) is used, the maximum LEO launch mass in the optimal mission architecture is 325 tons, and the total LEO mass of the system is 1142 tons. With the usage of NTP (specific impulse is 900 s) and NEP (the specific impulse is 6000 s) technology, the maximum LEO launch mass in the optimal mission architecture is only 85 tons, and the total LEO mass of the system is only about 400 tons. Considering the current rocket technology, the total cost is about USD 1149 million US.

1. Introduction

1.1. Manned Mars Exploration: Setting Sail

As one of the planets closest to Earth, Mars has long been regarded as an extremely important goal in the process of human exploration of the universe.
After a series of spacecraft visited Mars [1,2,3,4,5,6,7,8,9], our astronauts are expected to land on Mars in the near future. Unlike manned near-Earth missions (including manned missions to the moon/NEO), manned Mars missions will greatly challenge our existing space technology in terms of mission duration and mass scale. The journey to Mars will consume a surprising amount of propellant and supplies, while landing/taking off on Mars and returning to Earth will also require more powerful (and heavier) landing vehicles.
Following the success of the Apollo mission to the moon, space agencies of several countries have been demonstrating plans for manned missions to Mars. The most direct method of manned Mars exploration (the basic profile) is to use chemical engines to implement several orbital shifts. The spacecraft is pushed from LEO to the Earth–Mars transfer orbit and accelerates into LMO when it reaches its rendezvous with Mars. When returning, the spacecraft accelerates from Mars into Mars–Earth transfer orbit and finally enters the atmosphere. Many researchers have estimated that the initial mass of the spacecraft in LEO reached thousands of tons in this profile. This is obviously difficult to achieve at present.
NASA proposed the NASA 90-day study [10] in the 1980s, aiming to use the framework of the Apollo mission to implement manned Mars exploration. The plan uses the opposition trajectory to shorten the mission time, and at the same time, the mission risk is low. However, a spacecraft with a large initial mass (thousands of tons) is required.
Clark et al. [11] proposed the straight arrow scheme based on the basic profile, which suggested collecting and producing part of the propellant for partial return on the surface of Mars, thereby reducing the initial mass of the spacecraft. Zubrin and Frenchs also proposed ‘Mars Direct’ and ‘Case for Mars’ [12,13], which need to collect and produce all the propellants for return on the surface of Mars. The concept of multiple functional modules and separate launches of modular spacecraft is explored in these two studies. NASA has also proposed a series of plans for manned Mars exploration missions [14,15,16,17,18].
In addition, Taraba et al. proposed a manned Mars exploration scheme based on the LMO space station and preliminarily estimated the mass scale and cost of the manned Mars exploration missions [19]. Salotti et al. first proposed a manned Mars exploration scheme that divides the return spacecraft into two parts, which can complete the mission requirements when carrying only three astronauts [20]. In another study of Salotti et al., a new roadmap for the preparation of the human mission to Mars based on the work of ISECG is proposed [21]. He also proposes using two Mars rovers to carry two astronauts to explore the surface of Mars [22]. Colombi et al. proposed a radiation environment warning system for manned Mars exploration missions, which can effectively ensure the safety of astronauts [23].
On the other hand, the development of propulsion technology has brought dawn to manned deep space exploration missions. At present, mature propulsion technologies include high specific impulse chemical propulsion technology, electric propulsion technology, and nuclear propulsion technology.
In terms of chemical propulsion, the propellant H2/O2 is widely used as the propellant with a high specific impulse of about 440 s [24]. This kind of engine can provide large thrust (1 KN∼100 kN) and can operate stably. It is suitable for manned deep space exploration missions [25,26].
Electric propulsion technology greatly reduces the propellant mass with extremely high propellant specific impulse, thereby significantly reducing the initial mass of the spacecraft [27,28,29]. At present, most electric propulsion engines are powered by solar panels, and their thrust is very small (mN magnitude) due to the low power. Thus, they can only be used for small probes.
The gradual maturity of nuclear electric propulsion (NEP) technology gives us the opportunity to use larger thrust engines for manned Mars exploration missions [30]. Konstantinov analyzed a 1000-day manned Mars exploration mission based on the use of nuclear power engines [31], reducing the initial mass of the spacecraft to 200 tons. Loeb et al. further used the Megawatt Type nuclear reactor and the clustered Ion Thruster RFIT-45 to design a technically feasible NEP device for a manned Mars mission with an impulse of about 6000 s [32]. Such megawatt nuclear reactors are similar to those carried by nuclear submarines, and their technical conditions are relatively mature.
In addition, the development of nuclear thermal propulsion (NTP) technology also provides more options for impulse maneuver. NTP can provide greater thrust with a higher specific impulse. At present, the mature nuclear fission thermal propulsion mainly focuses on the solid core type, which can provide 100–1000 kN thrust, and the specific impulse can reach about 1000 s [33,34,35,36]. This technology can be used as an alternative to chemical engines in manned Mars exploration missions.
In terms of spacecraft design, the Boeing Company proposed a design plan for the Spaceship Discovery in 2006 [37]. There are two types of landing modules used by the Discovery spacecraft for Mars exploration. One is a manned landing module called Landing Module No.2 (LM2); the other is an automatically controlled cargo landing model, which is called Landing Module No.3 (LM3) [38,39]. Boden et al. also gave the design scheme of manned spacecraft in the research of manned asteroid missions [40]. This provides viable vehicle systems for manned Mars missions.

1.2. Mission Architecture Analysis: The Mission Architecture Matrix (MAM) Method

Since the manned Mars exploration vehicle system is very complex, it is necessary to conduct a modular mission scale analysis. The Mission Architecture Matrix (MAM) method is a modular mission analysis method based on a series of studies of baseline mission design of human deep space exploration [41,42,43,44,45]. This method transforms the mission flow into Basic Operation Steps (BOS) according to the baseline design. Each BOS corresponds to a Basic State Vector (BSV) composed of the state values of all modules of the system. The Basic Step Matrix (BSM) can then be constructed based on the corresponding BOS with BSV and related velocity increment, specific impulse, and other mission parameters. The BSM is actually the parameter transfer matrix of the system, connecting the mission parameters (such as mass, reliability, etc.) before and after one BOS. The parameter transition matrix at any two moments can be obtained by matrix operation of all BSMs between the two moments. Therefore, this method can concisely and intuitively describe the state of the system at all moments, including the initial and the final moment. The fuel consumption can also be obtained from the matrix relationship between the initial and final states. The schematic diagram of the MAM method is shown in Figure 1.
This method has been detailed, introduced, and successfully used for manned Lunar/ near-Earth asteroid mission analysis in our previous work [46,47]. This paper presents several mission architectures of manned Mars missions and then uses the MAM method to analyze the corresponding mission scale in Section 4. Both chemical engines and nuclear propulsion engines (NTP and NEP) will be included.

2. Mission Profile Calculation and Spacecraft System Design

Taking into account the revolution period of Mars and Earth, a whole mission takes about 3 years. Considering the survival status of astronauts, the manned Mars mission time is limited to 1000 days, and the residence time on Mars is no less than 300 days. The launch time takes place 2035–2050. The spacecraft system must include at least the command module and the Mars Landing module (including the ascending rocket). The mission profiles are calculated based on the premise that all spacecraft modules are first launched to LEO.

2.1. Mission Profiles Calculation for Impulse Maneuver Transfers

2.1.1. Hohmann Transfer

For spacecraft using impulse engines (chemical engines and NTP engines are considered in this paper), we can use the Hohmann transfer to accomplish the Mars exploration trajectory as Figure 2 shows.
Through the Hohmann transfer method, the velocity increments Δ v 1 Δ v 4 in the heliocentric coordinate system could be calculated referring to the work of Howard D. Curtis [48].
Note that, considering departure from LEO, in order to escape the gravitational pull of Earth, the spacecraft needs to travel a hyperbolic trajectory relative to Earth, arriving at its sphere of influence with a relative velocity greater than zero, as Figure 3 shows [49]. As for velocity transformation into the sphere of Influence of Earth (reentry to Earth), the spacecraft will cross the sphere of influence behind Earth and then impact its atmosphere from the hyperbolic approach trajectory.
Thus, actual Δ v 1 and Δ v 4 for orbit maneuver must be conversed according to v 1 and v 4 [48,49]:
Δ v 1 = v 1 2 + 2 v L E O 2 v L E O
Δ v 4 = v 4 2 + 2 v E I 2
In which v L E O is the orbit velocity of LEO (≈ 7.7 / s ), and v E I is the orbit velocity at the Earth Entry Interface (EI, 130 ) (≈ 7.8 / s ) [50].
Similarly, actual Δ v 2 and Δ v 3 for orbit maneuver from LMO can be derived:
Δ v 2 = v 2 2 + 2 v L M O 2 v L M O
Δ v 3 = v 3 2 + 2 v L M O 2 v L M O
In which v L M O is the orbit velocity of LMO (≈3.5 km/s).
The total impulse Δ v includes Δ v 1 Δ v 3 .
Calculated results of parameters of the Hohmann transfer for Earth–Mars (including mission duration and velocity increments) are listed in Table 1.
As seen in Table 1, the Hohmann transfer is indeed the most energy-efficient two-impulse maneuver for transferring between Earth and Mars or other two coplanar circular orbits sharing a common focus. However, it comes at the cost of excessive time consumption.
On the other hand, in the real situation, Mars and Earth orbits are not perfect circles. Consequently, there are a few differences in the transfer orbits in each rendezvous period, and the Hohmann transfer cannot accurately describe the Earth–Mars transfer opportunities with optimal energy in each rendezvous period. At this point, it comes down to solving Lambert’s problem.

2.1.2. Lambert Transfer

Concerned with the determination of an orbit from two position vectors and the time of flight, Lambert’s problem has important applications in the areas of rendezvous, targeting, guidance, and preliminary orbit determination. Thus, we can calculate the mission profiles and search their best launch/return windows based on the Lambert transfers.
Since the whole mission includes two Lambert transfers (Figure 4), it is necessary to calculate the velocity increments required for a series of Earth–Mars and Mars–Earth transfer orbits during 2035–2050, respectively, by providing the positions of Earth and Mars as well as the transfer times, and then selecting the appropriate launch window and return window according to the time constraints (mission time < 1000 days). Figure 5 shows the Lambert transfer velocity increment from Earth to Mars from 2035 to 2050, and Figure 6 shows the Lambert transfer velocity increment from Mars to earth from 2035 to 2050. The calculation method of Lambert transfer refers to the work of Vallado [51].
The feasible mission profiles correspond to the minimum total transfer velocity increments and meet the time constraints. Therefore, the mission profiles are decided by four time variables (launch window t 0 , transfer time from Earth to Mars T 1 = t 1 t 0 , return window t 2 , and transfer time from Mars to Earth T 2 = t 3 t 2 ):
V 1 , V 2 = L a m b e r t t 0 , t 1 , O r b i t E a r t h , O r b i t M a r s V 3 , V 4 = L a m b e r t t 2 , t 3 , O r b i t E a r t h , O r b i t M a r s t 3 t 0 1000 t 2 t 1 300
This optimization problem (minimum Δ v can be solved using the Particle swarm optimization (PSO) method (searching optimal launch window in each local optimal region) [52]. The optimal profiles (with minimum local velocity increments) are shown in Table 2. Analogously, all actual Δ v for orbit maneuver are conversed according to v .
As can be seen in Table 2, the mission profile with the smallest total impulse is the bold line in this table, which can be considered an alternative mission profile for manned Mars mission.

2.1.3. Multi-Impulse Transfer

In addition to the two-impulse transfers, there are also multi-impulse transfers available for interplanetary trajectories, which may help reduce the velocity increments required for a mission so as to realize the minimal energy target.
Suppose there are n successive Lambert transfers that form an n + 1 impulse transfer for a manned Mars mission. Let the initial conditions of the spacecraft for multi-impulse transfer be r 0 , v 0 , and t 0 , and the terminal conditions are r f , v f , and t f , in which r is the position vector of the spacecraft, and v is the velocity vector of the spacecraft. Assume that an impulsive Δ v i is applied; the conditions of the spacecraft before and after such impulse must satisfy
r i + = r i v i + = v i + Δ v i t i + = t i
In which i = 0 , 1 , , n .
Meanwhile, from the spacecraft orbital dynamic equation
r ¨ = μ r r 3
In which r is the modulus of r . We can solve to get
r i = f 1 r i 1 , v i 1 + , t i 1 , t i v i = f 2 r i 1 , v i 1 + , t i 1 , t i
In which i = 0 , 1 , , n .
According to the final conditions, using the above equations, we can obtain
r f = f 1 r n 1 , v n 1 + , t n 1 , t f v f = v n + Δ v n = f 2 r n 1 , v n 1 + , t n 1 , t f + Δ v n
Note that the transfer in every period between two consecutive impulses is the Lambert transfer, so we can obtain
v i 1 + , v i = Lambert r i 1 , r i , t i 1 , t i
In which i = 0 , 1 , , n .
In summary, take three-impulse transfer as an instance. To minimize the total velocity increments, the following constraints must be satisfied for the manned Mars mission:
min J = i = 1 n Δ v i s . t . v i + = v i + Δ v i r j = f 1 r j 1 , v j 1 + , t j 1 , t j v j = f 2 r j 1 , v j 1 + , t j 1 , t j r f = f 1 r 1 , v 1 + , t 1 , t f v f = f 2 r 1 , v 1 + , t 1 , t f + Δ v 2 v j 1 + , v j = Lambert r j 1 , r j , t j 1 , t j
In which i = 0 ∼2, j = 1 ∼2.
Similar to two-impulse transfer using only one Lambert transfer, the optimization problem (minimum Δ v ) can also be solved using the Particle swarm optimization (PSO) method by controlling the true anomaly θ or the transfer time Δ t of each period between two consecutive impulses or another variable.
Meanwhile, as some researchers have pointed out, for the coplanar cases, the use of a minimum delta-V three-impulse transfer affords scant improvement in velocity penalty [53]. Thus, considering the reliability, it seems that two-impulse transfers are better than multi-transfers for manned Mars missions.

2.1.4. Summary

By comparing all the transfer trajectories from Section 2.1.2 and Section 2.1.3 together, in this study, the mission profile with the smallest total impulse is taken for research; that is, the bold line in Table 2. The launch date of the mission is 5 September 2037, and the arrival date at Mars is 21 April 2038. After 453 days’ stay on Mars, the crew leaves Mars on 18 July 2039, and finally returns back to Earth on 15 May 2040. The total duration of the mission is about 983 days.

2.2. Mission Profiles Calculation for Low-Thrust Transfers

In this paper, only electric engines are considered for low-thrust transfers.
While using low-thrust transfers, there are some characteristics of manned Mars missions which need to be taken into consideration as mission constraints, such as the launch date, the duration of transfer to Mars, the program of spacecraft transfer Earth–Mars and Mars–Earth, the duration of stay of spacecraft on Martian orbit (minimal time is fixed 30 days), as well as the launching costs, the vehicle initial mass and the risk of a mission failure, in which the main performance index is the mass of nuclear electric power and propulsion system [31,32].
Based on the above-listed mission constraints, article [31] established some constraints for manned Mars missions using low-thrust transfers: the duration of the crew stay in space is 1000 days; the power of a jet flow of the electric propulsion is 1853.457 kW; the specific impulse of electric propulsion is 4500 s to 7500 s; the electric propulsion thrust is 50.4N to 84N; the specific mass of the system of Xenon storage (ratio of the mass of fuel compartment to Xenon mass a t a n k ) is equal to 0.1.
Furthermore, article [31] showed that at considered characteristics of a space transport system, the optimal magnitude of a specific impulse is equal to 7000 s. At an efficiency of electric propulsion 0.6, the specific mass of electric power and propulsion systems should not exceed 14.6 kg/kW. If the efficiency of electric propulsion is equal to 0.7, the specific mass of electric power and propulsion systems should not exceed 17.0 kg/kW.
Thus, as mentioned above, this research uses a nuclear power engine with a constant thrust of about 60 and a specific impulse of about 6000 s , which can generate a thrust acceleration of about 3   ×   10 4 m/s2 for a 200 tons spacecraft (the mass is estimated according to the analysis in Section 3).
Note that electric engines generally work at their rated power. The bang–bang control theory based on Pontryagin’s maximum principle is widely used in this type of small thrust trajectory design [54,55,56].
To solve optimal bang–bang control for low-thrust orbital transfers considering the first-order necessary optimality conditions derived from Lawden’s primer vector theory, Chen Zhang and Yu-Shan Zhao presented an original, hybrid switching detection technique [54]. T. Haberkorn et al. used a homotopic approach combined with the single shooting method (based on Pontryagin’s maximum principle), which transforms an optimal control problem by solving an equation of the shooting function associated with the original problem [55]. Zhengfan Zhu et al. developed a novel continuation technique [56] and successfully found fuel-optimal orbital transfers with variable mission parameters via an automated algorithm.
In this research, the low-thrust mission profile is calculated according to David Eagle’s work [57], which optimizes the steering angle with a fixed thrust. Figure 7 shows one of the optimal trajectories 2035–2050 based on the 60 N nuclear thruster in the heliocentric coordinate system. Figure 8 shows the corresponding steering angle function. The parameter information of the entire mission profile (Earth-to-Mars and Mars-to-Earth) is listed in Table 3. The spacecraft is only driven by low-thrust engines outside the influence sphere of central planets, so impulse maneuvers (provided by upper stage rockets or orbit transfer stages) are still used to escape from LEO/LMO ( 3.17 km/s for LEO and 1.45 km/s for LMO).

3. Spacecraft System Design

To reduce the mass of a single launch while ensuring the overall function of the spacecraft system, the modular spacecraft system design has been widely used in manned deep space exploration missions [58].
The spacecraft system in the manned Mars mission at least includes a manned module (command module) and a Mars lander. Extra propulsion modules, and sometimes an equipment module are also considered in the design of the mission.
B.D. Boden described a similar baseline spacecraft system including six modules for a manned Near Earth Objects (NEO) mission, but not considered a lander as NEOs always have a low-gravity environment [40]. A more complete spacecraft system for manned asteroid missions is concluded in the previous work [47] based on the modular spacecraft design method.
According to this method, spacecraft modules for manned Mars mission are then defined:
  • The Crew Exploration Vehicle (CEV). It mainly consists of a Command Module (CM) and Habitation Module (HM). Sometimes, a Propulsion Module (PM) is also needed for orbit and attitude adjustment. The mass of CEV with a crew of four people is about 60 tons (full of supply).
  • The Mars Lander (ML). This module may be similar to the Lander of Apollo, which contains a Descending Module (DM), and a separable Ascending Module (AM). Considering that the astronauts stay on the surface of Mars for a long time, a Long-term Mars Survival Cabin (MSC) is also necessary. The total mass of the Mars Lander is about 85 tons according to related research [31,32].
  • The chemical propulsion modules, which are always referred to as Orbit Transfer Stages (OTS), can provide velocity increments for the manned module(s). One OTS usually has a dry mass ratio which is proportional to its fueled mass ( η = m f / m 0 ; η is usually taken as 0.1). The most efficient specific impulse of OTS can reach about 440 s (liquid hydrogen–liquid oxygen engine, [24]).
  • The Nuclear Thermal Propulsion (NTP) module is used to provide impulse maneuver. According to the data of solid core NFTP motor RD-0410 in [36], the dry mass of the engine (reaction part) is about 2 tons. The specific impulse of the NTP module is approximated as 900 s. The dry mass ratio η is still 10 % .
  • The Nuclear Electric Propulsion (NEP) module is used to accelerate modules in the interplanetary journal. According to Leob’s research (Ion Thruster RFIT-45 with nuclear reactor), the dry mass of NEP is about 60 tons, and the specific impulse can reach about 6000 s [32].
  • The Supply Module (SM). Due to the long mission duration to Mars, astronauts may need additional supplies to survive. The SM mainly carries food, water, air, and other essential supplies. It can be delivered to Mars separately in advance and docks with CEV when needed. Considering that the SM can provide replenishment for a long mission time after CEV reaches Mars, it can replace about 20t mass of CEV (i.e., a 60 tons CEV can be replaced by a 40 tons CEV together with a 20 tons SM).
The basic parameters of spacecraft modules are listed in Table 4.

4. Mission Architectures Design and Mass Scale Analysis

This part will describe several mission architectures using chemical engines and NEP engines. The mission profiles (mainly transfer orbits and related delta-vs) in Section 2.1 and Section 2.2 are considered. For impulse maneuver engines (chemical engines and NTP engines), the launch date is 5 September 2037 (the second profile in Table 2, which has the smallest total velocity increment). For low-thrust (NEP) engines, the launch date is 7 June 2037.
There are four mission ‘nodes’ in a manned Mars mission including the Earth surface, the LEO, the LMO, and the Mars surface, thus dividing the manned Mars mission into seven successive mission ’phases’ (Figure 9). Therefore, the mission architectures are designed based on these four nodes and seven phases.

4.1. Architecture-1: The Basic Architecture

As described in Section 1.1, the basic profile for a manned Mars mission is a simple single process (similar to the Apollo mission), which contains launching from Earth, reaching LEO, accelerating twice to Mars ( Δ v 1 + Δ v 2 ), orbiting Mars on LMO, landing on Mars using ML, ML leaving Mars and back to LMO, accelerating from LEO to Earth and eventually re-entering the atmosphere. All spacecraft modules in this architecture are docked as one whole system at LEO and head to Mars together. The ML can be abandoned when returning from LMO to save propellant. The schematic of Architecture-1 is shown in Figure 10.
According to the MAM calculation method in [47], the Mission Architecture Matrix can be listed according to the basic operation steps (BOS) of mission architecture and the corresponding parameters. Each BOS is used as a row vector in the MAM, and the elements of BOS are BOS number, RVD flag (rendezvous and docking flag; one represents yes, whereas zero represents no), the status value of each module (zero for not participating, one for participating but not providing velocity increment, and two for providing velocity increment), the number (IMV) of the module which provides velocity increment, and the velocity increment corresponding to this BOS. Table 5 shows the MAM corresponding to Architecture-1.
Each BOS vector can be extended to an N × N state transition matrix (N is the number of modules), which is called the Basic Step Matrix (BSM). This matrix is used to connect the mass vectors (mass state of all modules) before and after BOS and generate them according to the following formula:
BSM = I N × N CMM · FCM
where I N × N is the n th order identity matrix. CMM (carrying module matrix) is an N × N matrix, elements in the participating row (row number is the number of the module with the value of one or two in the BOS vector) in the active column (corresponding to the active module, the column number is IMV) is one, and the other elements are zero. FCM (fuel calculation matrix) is an N × N matrix, in which a diagonal element corresponds to the active module (both row and column numbers are IMV). The value is:
FCM I M V , I M V = 1 e Δ v g I s p
and the other elements are zero. In Formula (13), Δ v is the velocity increment corresponding to this BOS, and I s p is the specific impulse corresponding to the active module. Take the second BOS of the MAM of Architecture-1 as an example, and its BSM is:
BSM = 1 1 1 0 1 0 1 1 · 0 0 1 e Δ v g I s p = 1 1 1 0 1 e Δ v g I s p 0 1 e Δ v g I s p 1 e Δ v g I s p = 1 e Δ v g I s p 1 1 e Δ v g I s p 1 e Δ v g I s p
Then, for the initial mass vector m 1 , m 2 , m 3 , there is:
m 1 , m 2 , m 3 = m 1 , m 2 , m 3 · BSM = m 1 , m 2 , m 3 · 1 e Δ v g I s p 1 1 e Δ v g I s p 1 e Δ v g I s p = m 1 , m 2 , ( m 1 + m 2 + m 3 ) · e Δ v g I s p ( m 1 + m 2 )
The final mass state exactly conforms to the Ziolkovsky formula.
Using the multiplicative property of matrix calculation, the Mass Transfer Matrix (MTM, which connects the initial and final mass status of the whole mission process) of Architecture-1, is:
MTM = i = 1 9 BSM i = 1 0.9 1 0.4 0.1
The mass of OTS can be then obtained:
M f = M 0 · MTM
where
M f = m f _ C E V , m f _ M L , m f _ O T S = 60000 , 85000 , m 0 _ O T S
M 0 = 60000 , 85000 , m 0 _ O T S
Formulas (16)∼(19) constitute matrix equations about m 0 _ O T S . When the dry mass ratio η is 0.1, the original mass of OTS is about 221 tons. Table 6 shows the mass scale of Architecture-1. Note, in Table 6, the minimum original mass has been bolded.
Obviously, Architecture-1 requires an LEO launch mass of up to 2200 tons, which is currently impossible for rockets. Therefore, mission architectures with more separate modules must be considered.

4.2. Architecture-2: Separate CEV and ML

Taking into account that ML only carries astronauts to perform landing operations after reaching Mars, CEV and ML can be launched to Mars separately as two parts, each carrying an OTS. This is described in Architecture-2. The schematic of Architecture-2 is shown in Figure 11. The mission architecture matrix of Architecture-2 is shown in Table 7.
Similarly to Formulas (16)∼(19), the mass of OTS1 and OTS2 can be obtained via matrix calculation.
Table 8 shows the mass scale of Architecture-2. Also, in Table 8, the minimum original mass has been bolded.
Although the maximum LEO launch mass required by Architecture-2 is still as high as 1700 tons, it has been reduced by about 23% compared to the basic profile (Architecture-1). This shows that the separate orbit transfer of different modules is an effective way to reduce the maximum LEO launch mass.

4.3. Architecture-3: Additional SM and Additional OTS for Return

It is noticed that the module with the largest LEO mass in Architecture-2 comes from OTS1, the CEV propulsion module. As we mentioned in Section 3, a 20-ton SM can be used to reduce the mass of CEV. At the same time, an OTS used for return acceleration can also be transferred to Mars along with SM. In this way, OTS1 is only used to transfer 40 tons of CEV to LMO. This is the profile of Architecture-3, which includes three independent transfer systems (CEV+OTS1, ML+OTS2, SM+OTS3+OTS4).
The schematic of Architecture-3 is shown in Figure 12. The mission architecture matrix of Architecture-3 is shown in Table 9. Based on the matrix calculation, Table 10 shows the mass scale of Architecture-3, in which the minimum original mass has been bolded.
The use of an additional supply module and returning propulsion module (OTS4) further reduced the maximum launch mass of LEO to 530 tons. At this time, the module with the largest launch mass is OTS2, which transports ML to Mars, while the LEO launch mass of OTS3, which carries the supply module and returning propulsion module, is also larger. We will further reduce the mass of these modules in Architecture-4.

4.4. Architecture-4: Two-Stage OTS for Earth to Mars Transfer

In fact, the main reason for the large mass of the transfer OTS is the large velocity increment required in the Earth-to-Mars transfer. Considering that there are mainly two impulse maneuvers in this process, a two-level OTS system can be used to complete two impulse maneuvers (dv1 and dv2), respectively. The first OTS can be discarded after the first pulse is completed, and the subsequent OCC and the second pulse are completed by the second OTS. Considering the two-level OTS method, Architecture-4 is obtained, which is shown in Figure 13, Table 11 and Table 12. In Table 12, the minimum original mass has been bolded.
In the case of a two-stage OTS design (one OTS for a single pulse), Architecture-4 is the profile with the lowest LEO mass requirement using chemical engines. At this time, we still need to launch an OTS weighing 325 tons to LEO, which is somewhat difficult with the current launch vehicle technology. The LEO carrying capacity of the Saturn V rocket with the largest thrust in human history is only about 140 tons. Therefore, it is necessary to use nuclear propulsion technology to send modules to Mars.

4.5. Architecture-5 8: Replace OTS with NTP Modules

According to the analysis in Section 3, NTP modules can actually be used in manned missions as OTS modules with higher specific impulses. Therefore, after replacing OTS in Architecture-1–4 with NTP without changing the basic architecture of the mission, the system mass scale can be significantly reduced. It is worth noting that the initial and final mass of one NTP module needs to meet:
m N T P , f = 2000   k g + η · ( m N T P , 0 m N T P , f )
where m N T P , f is the final mass, m N T P , 0 is the initial mass, η is the dry mass ratio of the NTP propellant system. 2000 kg is the dry weight of the nuclear engine.
Therefore, the mass scale of Architecture-5–8 corresponding to Architecture-1–4 is shown in Table 13, Table 14, Table 15 and Table 16. Note, in Table 13, Table 14, Table 15 and Table 16, each minimum original mass has been bolded.
After using NTP instead of OTS, the overall mass scale of the mission decreased significantly. When the mission architecture similar to Architecture-4 is adopted (Figure 13, Table 11), the maximum LEO launch mass is reduced to 85 tons (the Mars Lander), which is expected to be realized in the next few decades.

4.6. Architecture-9: The Basic Architecture Using NTP + NEP

Although the maximum LEO launch mass can be reduced to about 84 tons by using NTP technology, there are still technical difficulties in such a heavy single LEO launch. In addition, the overall mass scale of the mission is still high, which will result in high costs. In fact, the simultaneous use of NTP and NEP technology can further reduce the mass scale and the cost of the mission [33].
According to the research of Konstantinov, the NEP module can be used on the manned Martian spacecraft, which is designed as a whole system. This profile is equivalent to replacing OTS in Architecture-1 with a NEP module. The high specific impulse characteristic of NEP will greatly reduce fuel consumption. In addition, the velocity increment of each transfer will be calculated according to the thrust impulse integral during the transfer. NTP modules mainly provide impulse maneuvers to escape or enter LEO/LMO (Section 2.2). Using this method, the profile of Architecture-9 uses both NTP and NEP is shown in Figure 14 and Table 17.
Considering that the dry mass of the NEP module is about 60 tons and its specific impulse is 6000 s, the relationship between the initial mass of NEP and its fuel mass (mp) is
m 0 _ N E P m p _ N E P = m f _ N E P = 60 t o n s
Using Formulas (20) and (21), the mass scale can be obtained in Table 18 (similarly, in Table 18, the minimum original mass has been bolded):
In Architecture-9, a 165-ton NEP module and a 262-ton NTP module are needed for the entire spacecraft system to complete the Mars exploration journey. Although it is still difficult to send such heavy modules to LEO, inspired by Architecture-4, we can send different module combinations to Mars separately.

4.7. Architecture-10: Separate Transfer Using NTP and NEP

Although NEP and NTP technologies have been applied in Architecture-9, the maximum single LEO launch mass is still difficult to achieve. Therefore, similar to the profile in Architecture-4/8, three module combinations can be transferred, respectively. It is noted that the only propulsion module that exceeds 50 tons in Architecture-8 is NTP-3, which sends ML to Mars. Therefore, we consider replacing it with NEP. Considering that ML can be launched to LMO for standby in advance, the launch time of the ML part can be advanced, and NEP can be used for LEO escape and LMO entry, which may take a long time due to its low thrust [31].
For the manned part (CEV), two orbital maneuvers at the time of Earth departure and Mars arrival are realized by using two NTP modules (NTP-1, NTP-2). For the supply part (SM+supply NTP for Mars-to-Earth transfer), two NTP modules are also used (NTP-3, NTP-4). For the LM part, it can gradually escape LEO or enter LMO through long-term orbit maneuver using NEP, and the interplanetary transfer is also conducted by the NEP module. The profile of Architecture-10 is shown in Figure 15 and Table 19.
According to the MAM of this architecture, the mass scale of each module can be obtained (Table 20, in which the minimum original mass has been bolded):
Compared with Architecture-8, which only uses NTP, the maximum LEO launch mass in the entire mission is still 85 tons (the Mars Lander), while the total LEO mass scale of Architecture-10 is about 8% lower than Architecture-8. On the other hand, Architecture-10 using NEP for low thrust transfer may require a more complex control system.

4.8. Total Cost

According to Aerospace Security’s statistics [59], considering the current rocket technology, the cost of a space launch to LEO is approximately USD 15,500 per 1 kg. As for spacecraft and launch vehicle system development and manufacturing, after comparing and analyzing the budgets of various programs, including the Artemis program, we estimate the cost to be approximately USD 500 million.
Hence, based on the above, the total cost of 10 mission architectures in Section 4.1, Section 4.2, Section 4.3, Section 4.4, Section 4.5, Section 4.6 and Section 4.7 can be calculated, as Figure 16 shows.
As can be seen from Figure 16, the total cost of 10 architectures approximately ranges from USD 1.1 billion to USD 4.2 billion. Among them, Architecture-10 enjoys the minimum total cost of USD 1.1 billion due to its low mass scale.

4.9. Summarize

We analyzed and calculated the mass scale of 10 mission architectures in Section 4.1, Section 4.2, Section 4.3, Section 4.4, Section 4.5, Section 4.6 and Section 4.7. The total mass scale and maximum LEO launch mass of 10 architectures are shown in Figure 17 and Figure 18.
As we can see from Figure 17 and Figure 18, the total mass scale of 10 architectures approximately ranges from 385 tons to 237 tons, while the maximum LEO launch mass ranges from 85 tons to about 2200 tons. Among them, the total mass scale of Architecture-10 is minimal, at around 385 tons, and the maximum LEO launch mass of Architecture-8 as well as Architecture-10 is minimal at 85 tons.
It is not difficult to find that the total mass scale and the maximum LEO launch mass of Architecture-1 and Architecture-2 are both excessively large and far beyond the other architectures due to their excessively low number of launches and modules. When using additional modules such as SM and OTS to reduce the transfer mass of every OTS, there exists a visible reduction in the mass scale, and the maximum LEO launch mass has dropped to 500 tons or below. Since the transfer mass of OTS is reduced, the required fuel mass is correspondingly reduced, resulting in an overall reduction in mission mass scale.
With the use of NTP and NEP, attributed to their higher specific impulse, the fuel mass required for the architecture is significantly reduced while the payload remains the same, greatly decreasing the total mass scale of Architecture-5–10 to less than 500 tons, except for Architecture-9 because of its Apollo style (a single launch and no additional module such as SM). As for the maximum LEO launch mass, the situation is similar.
Generally, the results show that the combined use of nuclear propulsion, the modular launch and separate transfer can greatly reduce the fuel mass required by the mission, to reduce the mass scale of the mission, as well as the maximum LEO launch mass.
Undoubtedly, the maximum LEO launch mass is a rather important evaluation factor for manned Mars mission architectures, as it involves the requirements for the carrying capacity of the launch vehicles. We can obtain the LEO carrying capacity (unit: tons ) of some large and heavy-lift launch vehicles (Figure 19) [60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76].
Let us take three heavy-lift launch vehicles under development, including Starship, SLS Block along with Long March 9, and compare their LEO carrying capacity to the maximum LEO launch mass of 10 architectures, as Figure 20 shows.
After comparison, it can be found that with the exception of Architecture-8 and Architecture-10, the maximum LEO launch mass of nearly all architectures exceeds the LEO carrying capacity of Starship. The SLS Block is slightly better, and can meet the requirements of Architecture-6, 7, 8 and 10. Long March 9 can meet the launch quality requirements of other architectures except Architecture-1, 2, 3.
Note that although NTP and NEP both have a significant advantage in efficiency over traditional chemical rocket technologies, the nuclear reactors in NTP and NEP are more complex and have higher individual costs.
Moreover, as emerging aerospace propulsion systems undergo development, the reliability of NTP and NEP is relatively insufficient. As of today, no nuclear propulsion system has been officially put into service.
As the number of launches and the number of modules (especially for nuclear modules) increases, the complexity of an architecture increases, which in turn reduces the reliability of the architecture.
Clearly, a coordinated trade-off between cost and reliability is necessary.
According to the maximum LEO launch mass constraint (<100 tons), the total cost and the launch vehicle capacity limitations, two feasible architectures are selected, namely Architecture-8 and Architecture-10. Considering the technical maturity and control complexity, Architecture-8 is the best choice. Figure 21 shows the detailed mission profile of Architecture-8.

5. A Feasible Mission Profile

After considering modular launch, separate transfer, and use of NTP/NEP modules, we have obtained a feasible mission architecture (Architecture-8) in Section 4.9 (Figure 21, Table 16). This mission architecture includes 10 modules (CEV, ML, SM, and 7 NTP modules), with a total LEO launch mass of 419 tons and a maximum LEO launch mass of 85 tons (for ML). In order to reduce the number of ground launches as much as possible, the launches are planned in Table 21 according to the mass scale estimated in Table 16, where the maximum launch mass cannot exceed 100 tons.
According to Table 21 and Section 4.8, it can be estimated that the total cost of a manned Mars mission is about USD 1149 million.
Combined with the analysis of the launch window in Section 2.1.2, Table 22 shows a detailed flow-chart of the mission profile for manned Mars exploration.

6. Conclusions and Discussion

Using the modular design of the spacecraft system and the NTP/NEP technology, manned Mars exploration missions are expected to be realized in the near future. This paper uses the MAM method to analyze the mass scale of several multiple-mission architectures. Finally, the mission architecture using the NTP (Nuclear Thermal Propulsion) engines and the modular transfer method was determined (Figure 21). The Crew Exploration Vehicle, the Mars Lander, and the supply combination (the Supply Module and the supply NTP stage) is sent to Mars separately using two-stage NTP modules. All modules dock on the LMO, and then the Mars landing mission is executed. The return spacecraft system (the Crew Exploration Vehicle and the Supply module) finally flies back to Earth using the supply NTP stage.
The total LEO mass scale of the optimal architecture is about 419 tons, and its maximum LEO launch mass is only about 85 tons, which can be achieved by multiple launches by heavy launch vehicles. The total launch cost is about USD 1149 million.
According to this mission architecture, we designed a specific manned mission flow-chart to Mars. The three module combinations will be sent from LEO into the Earth–Mars transfer orbit on 5 September 2037, separately. The crew will stay on Mars from 21 April 2038, to 18 July 2039, and will return to Earth on 15 May 2040. The total mission duration is 983 days.
Throughout this paper, it is reasonable to assume that, compared with traditional research methods, the Mission Architecture Matrix method has the advantages of high efficiency and accuracy of calculation and analysis, conciseness, and strong flexibility. The method can be applied not only to the analysis of manned Mars missions, but also to the analysis of other space missions, including complex deep space exploration.
As seen in Section 4, the use of nuclear propulsion technology is a good option for manned deep space exploration missions with large mass scale and large velocity increments.
Of course, it must be acknowledged that nuclear propulsion technology is still immature. At present, as far as international research is concerned, there is more research on NEP technology, and NASA and Russia and so on have made progress. Regarding NTP, although related research has long been conducted to design and analyze a series of nuclear thermal propulsion engines, ranging from the solid reactor up to the gas core reactor, there still exist numerous problems and challenges to be solved. All early applications for NTP used fission processes; only in the last decade has there been a shift to fusion approaches.
As emerging aerospace propulsion systems undergo development, nuclear propulsion technology (especially for NTP) faces some risks, such as complex nuclear reactors, the dispersal of radioactive material into the environment resulting from an atmospheric or orbital rocket failure, the impact of cosmic ray particles on nuclear energy engines, and so forth, which needs further research in the future. We can believe that with the solution to these problems, nuclear propulsion technology will propel mankind into the vast universe.

Author Contributions

Conceptualization, X.W. and Y.Y.; methodology, D.S.; software, D.S. and Y.Y.; validation, D.S., X.W. and Y.Y.; formal analysis, D.S.; investigation, D.S.; resources, Y.Y.; data curation, D.S.; writing—original draft preparation, D.S.; writing—review and editing, D.S. and Y.Y.; visualization, D.S.; supervision, X.W.; project administration, X.W.; funding acquisition, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by National Natural Science Foundation of China (No. 41804165).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Branigan, T.L. Mariner 4: Mission to mars. Phys. Teach. 1965, 3, 303–307. [Google Scholar] [CrossRef]
  2. Siddiqi, A.A. Beyond Earth: A Chronicle of Deep Space Exploration, 1958–2016; National Aeronautics and Space Administration, Office of Communications: Washington, DC, USA, 2018. [Google Scholar]
  3. Cooley, C. Viking 75 Project: Viking Lander System Primary Mission Performance Report; Technical Report; Martin Marietta Corp.: Baltimore, MD, USA, 1977. [Google Scholar]
  4. Bickler, D. Roving over mars. Mech. Eng. 1998, 120, 74–77. [Google Scholar] [CrossRef]
  5. Graf, J.E.; Zurek, R.W.; Erickson, J.K.; Jai, B.; Johnston, M.; de Paula, R. Status of Mars reconnaissance orbiter mission. Acta Astronaut. 2007, 61, 44–51. [Google Scholar] [CrossRef]
  6. Zurek, R.W.; Smrekar, S.E. An overview of the Mars Reconnaissance Orbiter (MRO) science mission. J. Geophys. Res. Planets 2007, 112, E5. [Google Scholar] [CrossRef] [Green Version]
  7. Saunders, R.; Arvidson, R.; Badhwar, G.; Boynton, W.; Christensen, P.; Cucinotta, F.; Feldman, W.; Gibbs, R.; Kloss, C.; Landano, M.; et al. 2001 Mars Odyssey mission summary. Space Sci. Rev. 2004, 110, 1–36. [Google Scholar] [CrossRef] [Green Version]
  8. Shotwell, R. Phoenix—The first Mars Scout mission. Acta Astronaut. 2005, 57, 121–134. [Google Scholar] [CrossRef]
  9. Li, C.; Cong, Z. China successfully launched its first mars mission, Tianwen 1. China Space News 2020, 21, 51. [Google Scholar]
  10. States, U. Report of the 90-Day Study on Human Exploration of the Moon and Mars; National Aeronautics and Space Administration: Washington, DC, USA, 1989. [Google Scholar]
  11. Clark, B.C. A straight-arrow approach for the near-term human exploration of Mars. In The Case for Mars IV: The International Exploration of Mars- Mission Strategy and Architectures; Published for the American Astronautical Society by Univelt: San Diego, CA, USA, 1997; p. 1997. [Google Scholar]
  12. Zubrin, R.; Baker, D.; Gwynne, O. Mars direct—A simple, robust, and cost effective architecture for the space exploration initiative. In Proceedings of the 29th Aerospace Sciences Meeting, Reno, NV, USA, 7–10 January 1991; p. 329. [Google Scholar]
  13. Welch, S. Mission strategy and spacecraft design for a Mars Base Program. In The Case for Mars II; Published for the American Astronautical Society by Univelt: San Diego, CA, USA, 1985; pp. 344–375. [Google Scholar]
  14. Portree, D.S. Humans to Mars: Fifty Years of Mission Planning, 1950–2000; Number 20; National Aeronautics and Space Administration: Washington, DC, USA, 2001. [Google Scholar]
  15. Drake, B.G.; Hoffman, S.J.; Beaty, D.W. Human exploration of Mars, design reference architecture 5.0. In Proceedings of the 2010 IEEE Aerospace Conference, Big Sky, MT, USA, 6–13 March 2010; pp. 1–24. [Google Scholar]
  16. Glaze, L.S.; Baloga, S.M.; Garry, W.B.; Fagents, S.A.; Parcheta, C. A hybrid model for leveed lava flows: Implications for eruption styles on Mars. J. Geophys. Res. Planets 2009, 114, E7. [Google Scholar] [CrossRef] [Green Version]
  17. National Aeronautics. NASA’s Journey to Mars: Pioneering Next Steps in Space Exploration; Government Printing Office: Washington, DC, USA, 2015. [Google Scholar]
  18. Giannone, M. A Solution for Medical Needs and Cramped Quarters in Space IVGEN Undergoes Lifetime Testing in Preparation for Future Missions. 2012. Available online: https://www.nasa.gov/mission_pages/station/research/news/IVGEN.html (accessed on 30 August 2022).
  19. Taraba, M.; Zwintz, K.; Bombardelli, C.; Lasue, J.; Rogler, P.; Ruelle, V.; Schlutz, J.; Schüßler, M.; O’Sullivan, S.; Sinzig, B.; et al. Project M—A study for a manned Mars mission in 2031. Acta Astronaut. 2006, 58, 88–104. [Google Scholar] [CrossRef]
  20. Salotti, J.M. Robust, affordable, semi-direct Mars mission. Acta Astronaut. 2016, 127, 235–248. [Google Scholar] [CrossRef]
  21. Salotti, J.M.; Heidmann, R. Roadmap to a human Mars mission. Acta Astronaut. 2014, 104, 558–564. [Google Scholar] [CrossRef]
  22. Salotti, J.M. Simplified scenario for manned Mars missions. Acta Astronaut. 2011, 69, 266–279. [Google Scholar] [CrossRef]
  23. Colombi, J.M.; Miller, M.E.; Bohren, J.S.; Howard, J.K. Conceptual design using executable architectures for a manned mission to Mars. IEEE Syst. J. 2014, 9, 495–507. [Google Scholar] [CrossRef]
  24. Larson, W.J.; Pranke, L.K. Human Spaceflight: Mission Analysis and Design; The McGraw-Hill Companies: New York, USA, 2007. [Google Scholar]
  25. Zhang, Y.; Fu, J.; Xie, M.; Liu, J. Improvement of H2/O2 chemical kinetic mechanism for high pressure combustion. Int. J. Hydrogen Energy 2021, 46, 5799–5811. [Google Scholar] [CrossRef]
  26. Limerick, C. Dual mixture ratio H2/O2 engine for single stage to orbit application. J. Propuls. Power 1991, 7, 31–36. [Google Scholar] [CrossRef]
  27. Saccoccia, G.; Berry, W. European electric propulsion activities and programmes. Acta Astronaut. 2000, 47, 193–203. [Google Scholar] [CrossRef]
  28. Saccoccia, G.; del Amo, G. Electric propulsion—A key technology for space missions in the new millennium. ESA Bull. 2000, 101, 62–71. [Google Scholar]
  29. Sankovic, J.M. NASA technology investments in electric propulsion: New directions in the new millennium. In Proceedings of the Third International Conference on Spacecraft Propulsion, Berlin/Heidelberg, Germany, 11–13 August 2002; number E-12569. [Google Scholar]
  30. McGinnis, S.J. Nuclear Power Systems for Manned Mission to Mars; Storming Media: Washington, DC, USA, 2004. [Google Scholar]
  31. Konstantinov, M.S.; Petukhov, V.G. The analysis of manned Mars mission with duration of 1000 days. Acta Astronaut. 2012, 73, 122–136. [Google Scholar] [CrossRef]
  32. Loeb, H.; Petukhov, V.; Popov, G.; Mogulkin, A. A realistic concept of a manned Mars mission with nuclear–electric propulsion. Acta Astronaut. 2015, 116, 299–306. [Google Scholar] [CrossRef]
  33. Gabrielli, R.A.; Herdrich, G. Review of nuclear thermal propulsion systems. Prog. Aerosp. Sci. 2015, 79, 92–113. [Google Scholar] [CrossRef]
  34. Walton, J. An overview of tested and analyzed NTP concepts. In Proceedings of the Conference on Advanced SEI Technologies, Cleveland, OH, USA, 4–6 September 1991; p. 3503. [Google Scholar]
  35. Kulcinski, G.L. Nuclear Power in Space. 2000. Available online: https://fti.neep.wisc.edu/fti.neep.wisc.edu/neep602/FALL97/LEC22/lecture22.html (accessed on 30 August 2022).
  36. KBKhA. RD0410. 2014. Available online: https://kbkha.ru/deyatel-nost/raketnye-dvigateli-ao-kbha/yadernyj-raketnyj-dvigatel-rd0410-rd0411/ (accessed on 30 August 2022).
  37. Benton, M. Spaceship Discovery-Space Vehicle System For Human Interplanetary Exploration. In Proceedings of the 42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Sacramento, CA, USA, 9–2 July 2006; p. 5128. [Google Scholar]
  38. Benton, M. Conceptual Design of Mars Crew and Cargo Exploration Landers for Spaceship Discovery. In Proceedings of the AIAA SPACE 2007 Conference & Exposition, Virtual, Long Beach, CA, USA, 18–20 September 2007; p. 6239. [Google Scholar]
  39. Benton Sr, M.G. Spaceship Discovery’s Crew and Cargo Lander Module Designs for Human Exploration of Mars. In AIP Conference Proceedings; American Institute of Physics: College Park, MD, USA, 2008; Volume 969, pp. 898–907. [Google Scholar]
  40. Boden, R.C.; Hein, A.M.; Kawaguchi, J. Target selection and mass estimation for manned NEO exploration using a baseline mission design. Acta Astronaut. 2015, 111, 198–221. [Google Scholar] [CrossRef]
  41. Hofstetter, W. Extensible Modular Landing Systems for Human Moon and Mars Exploration. Master’s Thesis, Institute of Astronautics, TU Munich, Munich, Germany, 2004. [Google Scholar]
  42. Thompson, R.W. System architecture modeling for technology portfolio management using atlas. In Proceedings of the 2006 IEEE Aerospace Conference, Big Sky, MT, USA, 4–11 March 2006; 9p. [Google Scholar]
  43. Herman, J.F.; Zimmer, A.K.; Reijneveld, J.P.; Dunlop, K.L.; Takahashi, Y.; Tardivel, S.; Scheeres, D.J. Human exploration of near earth asteroids: Mission analysis for chemical and electric propulsion. Acta Astronaut. 2014, 104, 313–323. [Google Scholar] [CrossRef]
  44. Simmons, W.; Koo, B.; Crawley, E. Architecture generation for Moon-Mars exploration using an executable meta-language. In Space 2005; American Institute of Aeronautics and Astronautics: Commonwealth of Virginia, USA, 2005; p. 6726. [Google Scholar]
  45. Bounova, G.; Ahn, J.; Hofstetter, W.; Wooster, P.; Hassan, R.; de Weck, O. Selection and technology evaluation of moon/mars transportation architectures. In Space 2005; American Institute of Aeronautics and Astronautics: Commonwealth of Virginia, USA, 2005; p. 6790. [Google Scholar]
  46. Wang, X.; Mao, L.; Yue, Y.; Zhao, J. Manned lunar landing mission scale analysis and flight scheme selection based on mission architecture matrix. Acta Astronaut. 2018, 152, 385–395. [Google Scholar] [CrossRef]
  47. Yue, Y.; Wang, X.; Gu, H.; Mao, L. Mission architecture analysis for manned NEA exploration using MAM method. Acta Astronaut. 2020, 168, 256–272. [Google Scholar] [CrossRef]
  48. Curtis, H. Orbital Mechanics for Engineering Students; Butterworth-Heinemann: Oxford, UK, 2013. [Google Scholar]
  49. Weber, B. Orbital Mechanics & Astrodynamics. 2021. Available online: https://orbital-mechanics.space/intro.html (accessed on 30 August 2022).
  50. Walter, U.; Walter, U. Astronautics; Springer: Berlin/Heidelberg, Germany, 2008; p. 167. [Google Scholar]
  51. Vallado, D.A. Fundamentals of Astrodynamics and Applications; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2001; Volume 12. [Google Scholar]
  52. Bonyadi, M.R.; Michalewicz, Z. Particle swarm optimization for single objective continuous space problems: A review. Evol. Comput. 2017, 25, 1–54. [Google Scholar] [CrossRef]
  53. Bean, W. Noncoplanar minimum DELTA V two-impulse and three-impulse orbital transfer from a regressing oblate earth assembly parking ellipse onto a flyby trans-Mars asymptotic velocity vector. Astronaut. Acta 1971, 16, 19720029968. [Google Scholar]
  54. Zhang, C.; Topputo, F.; Bernelli-Zazzera, F.; Zhao, Y.S. Low-thrust minimum-fuel optimization in the circular restricted three-body problem. J. Guid. Control. Dyn. 2015, 38, 1501–1510. [Google Scholar] [CrossRef] [Green Version]
  55. Haberkorn, T.; Martinon, P.; Gergaud, J. Low thrust minimum-fuel orbital transfer: A homotopic approach. J. Guid. Control. Dyn. 2004, 27, 1046–1060. [Google Scholar] [CrossRef] [Green Version]
  56. Zhu, Z.; Gan, Q.; Yang, X.; Gao, Y. Solving fuel-optimal low-thrust orbital transfers with bang-bang control using a novel continuation technique. Acta Astronaut. 2017, 137, 98–113. [Google Scholar] [CrossRef]
  57. David Eagle. Low-thrust Earth-to-Mars Trajectory Analysis—SNOPT. 2021. Available online: https://www.mathworks.com/matlabcentral/fileexchange/42374-low-thrust-earth-to-mars-trajectory-analysis-snopt (accessed on 30 August 2022).
  58. Sforza, P.M. Manned Spacecraft Design Principles; Elsevier: Amsterdam, The Netherlands, 2015. [Google Scholar]
  59. Thomas, G.; Roberts. Space Launch to Low Earth Orbit: How Much Does It Cost? 2022. Available online: https://aerospace.csis.org/data/space-launch-to-low-earth-orbit-how-much-does-it-cost/ (accessed on 30 August 2022).
  60. Arianespace. ARIANE 5: The Heavy Launcher. 2022. Available online: https://www.arianespace.com/vehicle/ariane-5/ (accessed on 30 August 2022).
  61. SpaceX. Falcon 9: First Orbital Class Rocket Capable of Reflight. 2022. Available online: https://www.spacex.com/vehicles/falcon-9/ (accessed on 30 August 2022).
  62. Mission, A.Y. Proton Launch System Mission Planner’s Guide Section 2 LV Performance; International Launch Services: Commonwealth of Virginia, VA, USA, 2009. [Google Scholar]
  63. Khrunichev SRPSC. Angara Launch Vehicles Family. 2022. Available online: http://www.khrunichev.ru/main.php?id=44&lang=en (accessed on 30 August 2022).
  64. CNSA. (CZ-5). 2017. Available online: http://www.cnsa.gov.cn/n6758824/n6759008/n6759011/c6794064/content.html (accessed on 30 August 2022).
  65. Alliance, U.L.; Delta IV Launch Services User’s Guide. June 2013. Available online: https://www.ulalaunch.com/docs/default-source/rockets/delta-iv-user’s-guide.pdf (accessed on 30 August 2022).
  66. SpaceX. Falcon Heavy: The World’S Most Powerful Rocket. 2022. Available online: https://www.spacex.com/vehicles/falcon-heavy/ (accessed on 30 August 2022).
  67. Arianespace. ARIANE 6: The Next-Generation Launch Vehicle. 2022. Available online: https://www.arianespace.com/ariane-6/ (accessed on 30 August 2022).
  68. Foust, J. Eutelsat First Customer for Blue Origin’s New Glenn. 2017. Available online: https://spacenews.com/eutelsat-first-customer-for-blue-origins-new-glenn/ (accessed on 7 March 2017).
  69. GlobalSecurity. CZ-5DY Long March 5DY—New Generation of Manned Launch Vehicle. 2021. Available online: https://www.globalsecurity.org/space/world/china/cz-5dy.htm (accessed on 4 November 2021).
  70. SpaceX. Starship. 2022. Available online: https://www.spacex.com/vehicles/starship/ (accessed on 30 August 2022).
  71. TASS. Russia to Launch Super-Heavy Rocket to Moon in 2032–2035. 2018. Available online: https://tass.com/science/986450 (accessed on 23 January 2018).
  72. Creech, S. NASA’s Space Launch System: A Capability for Deep Space Exploration; NASA: Washington, DC, USA, 2014. [Google Scholar]
  73. Qin, X.; Long, L.; Rong, Y. The achievement and future of China space transportation system. J. Deep Space Explor. 2016, 3, 315–322. [Google Scholar]
  74. Energia. RSC “Energia”—History. 2022. Available online: https://www.energia.ru/en/history/systems/vehicles/vehicle_n1-l3_c.html (accessed on 30 August 2022).
  75. Energia, S.P. Korolev RSC Energia—LAUNCHERS. 2022. Available online: https://www.energia.ru/english/energia/launchers/vehicle_energia.html (accessed on 30 August 2022).
  76. Rehmus, P.B. Alternatives for Future US Space-launch Capabilities; Congress of the US, Congressional Budget Office: Washington, DC, USA, 2006. [Google Scholar]
Figure 1. Schematic diagram of the MAM method.
Figure 1. Schematic diagram of the MAM method.
Aerospace 09 00604 g001
Figure 2. Mission profile for impulse maneuver transfers (Hohmann transfers) in heliocentric coordinate system.
Figure 2. Mission profile for impulse maneuver transfers (Hohmann transfers) in heliocentric coordinate system.
Aerospace 09 00604 g002
Figure 3. Departure of a spacecraft on a mission from an inner planet to an outer planet.
Figure 3. Departure of a spacecraft on a mission from an inner planet to an outer planet.
Aerospace 09 00604 g003
Figure 4. Mission profile for impulse maneuver transfers (Lambert transfers) in heliocentric coordinate system.
Figure 4. Mission profile for impulse maneuver transfers (Lambert transfers) in heliocentric coordinate system.
Aerospace 09 00604 g004
Figure 5. Total Δ v for Earth-to-Mars Lambert transfer from 2035 to 2050. The horizontal axis is launch windows (in days after 1 January 2035), and the vertical axis is transfer time (in days).
Figure 5. Total Δ v for Earth-to-Mars Lambert transfer from 2035 to 2050. The horizontal axis is launch windows (in days after 1 January 2035), and the vertical axis is transfer time (in days).
Aerospace 09 00604 g005
Figure 6. Total Δ v for Mars-to-Earth Lambert transfer from 2035 to 2050. The horizontal axis is launch windows (in days after 1 January 2035), and the vertical axis is transfer time (in days).
Figure 6. Total Δ v for Mars-to-Earth Lambert transfer from 2035 to 2050. The horizontal axis is launch windows (in days after 1 January 2035), and the vertical axis is transfer time (in days).
Aerospace 09 00604 g006
Figure 7. Heliocentric trajectories Earth to Mars (units in meters). This is one of the optimal mission profiles based on low-thrust orbital transfers using electric propulsion engines. The sun is at centre of the coordinate, the blue orbit is the Earth orbit, and the orange orbit is the Mars orbit. The black curve is the trajectory of the transfer.
Figure 7. Heliocentric trajectories Earth to Mars (units in meters). This is one of the optimal mission profiles based on low-thrust orbital transfers using electric propulsion engines. The sun is at centre of the coordinate, the blue orbit is the Earth orbit, and the orange orbit is the Mars orbit. The black curve is the trajectory of the transfer.
Aerospace 09 00604 g007
Figure 8. Steering angle (deg) for the mission profile in Figure 7. The horizontal axis is transfer time (in days).
Figure 8. Steering angle (deg) for the mission profile in Figure 7. The horizontal axis is transfer time (in days).
Aerospace 09 00604 g008
Figure 9. Four mission nodes and seven mission phases in manned Mars mission.
Figure 9. Four mission nodes and seven mission phases in manned Mars mission.
Aerospace 09 00604 g009
Figure 10. Schematic of Architecture-1.
Figure 10. Schematic of Architecture-1.
Aerospace 09 00604 g010
Figure 11. Schematic of Architecture-2.
Figure 11. Schematic of Architecture-2.
Aerospace 09 00604 g011
Figure 12. Schematic of Architecture-3.
Figure 12. Schematic of Architecture-3.
Aerospace 09 00604 g012
Figure 13. Schematic of Architecture-4.
Figure 13. Schematic of Architecture-4.
Aerospace 09 00604 g013
Figure 14. Schematic of Architecture-9.
Figure 14. Schematic of Architecture-9.
Aerospace 09 00604 g014
Figure 15. Schematic of Architecture-10.
Figure 15. Schematic of Architecture-10.
Aerospace 09 00604 g015
Figure 16. Total cost of 10 different architecture.
Figure 16. Total cost of 10 different architecture.
Aerospace 09 00604 g016
Figure 17. Total mass scale of 10 different architecture.
Figure 17. Total mass scale of 10 different architecture.
Aerospace 09 00604 g017
Figure 18. Maximum LEO launch mass of 10 different architecture.
Figure 18. Maximum LEO launch mass of 10 different architecture.
Aerospace 09 00604 g018
Figure 19. Current LEO carrying capacity of large and heavy-lift launch vehicles.
Figure 19. Current LEO carrying capacity of large and heavy-lift launch vehicles.
Aerospace 09 00604 g019
Figure 20. Comparison of the maximum LEO launch mass of 10 architectures with the LEO carrying capacity of selected launch vehicles.
Figure 20. Comparison of the maximum LEO launch mass of 10 architectures with the LEO carrying capacity of selected launch vehicles.
Aerospace 09 00604 g020
Figure 21. Mission profile of Architecture-8.
Figure 21. Mission profile of Architecture-8.
Aerospace 09 00604 g021
Table 1. Parameters ofthe Hohmann transfer for Earth–Mars and Mars–Earth. The velocity increments in this table are all actual Δ v departure from LEO/LMO.
Table 1. Parameters ofthe Hohmann transfer for Earth–Mars and Mars–Earth. The velocity increments in this table are all actual Δ v departure from LEO/LMO.
Mission Duration (Days)Velocity Increments (km/s)
Total DurationEarth-MarsMars ResidenceMars-EarthTotal Impulse Δ v Δ v 1 Δ v 2 Δ v 3 Δ v 4
972.1258.9454.4258.97.83.62.12.111.5
Table 2. Optimal mission profiles for manned Mars mission using chemical engines. Time of expedition: <1000 days. The launch window is 2035–2050. The velocity increments in this table are all actual Δ v departure from LEO/LMO.
Table 2. Optimal mission profiles for manned Mars mission using chemical engines. Time of expedition: <1000 days. The launch window is 2035–2050. The velocity increments in this table are all actual Δ v departure from LEO/LMO.
Launch
Date
(dd.mm.yy)
Mission Duration (Days)Velocity Increments (km/s)
Total
Duration
Earth–MarsMars ResidenceMars–EarthTotal
Impulse Δ v
Δ v 1 Δ v 2 Δ v 3 Δ v 4
09.07.2035928.2184.3544.1199.78.43.62.12.711.9
05.09.2037982.9227.6453.1302.28.44.02.02.311.5
14.10.2039974.0224.2433.1316.88.64.02.42.211.6
24.11.2041977.7208.7431.2337.99.24.23.02.011.6
03.01.2044987.3185.6481.2320.69.94.23.72.012.1
12.02.2046978.5155.8546.7276.110.04.33.81.912.0
05.04.2048988.4151.1598.4238.98.73.93.01.911.6
06.06.2050977.4161.8589.8225.88.53.62.52.411.6
Table 3. Parameters for one optimal mission profile for manned Mars mission using electric propulsion engines. Time of expedition: <1000 days. The launch window is 2035–2050.
Table 3. Parameters for one optimal mission profile for manned Mars mission using electric propulsion engines. Time of expedition: <1000 days. The launch window is 2035–2050.
Transfer DurationStart TimeEnd TimeTotal Δ v  1
Earth-to-Mars300 days7 June 20373 April 20386.1 km/s
Mars-to-Earth300 days11 August 20396 June 20409.4 km/s
1 The total velocity increment is obtained by the propellant mass.
Table 4. The basic parameters of all considered modules.
Table 4. The basic parameters of all considered modules.
ModulesCEVMLOTSNTPNEPSM
Dry mass (tons)6085- 2 + x N T P 16020
Dry mass ratio--0.1--
Specific impulse--440 s900 s6000 s-
1 The dry mass of NTP consists of the dry mass of nuclear engine (2 tons) and the propellant storage system (which can be estimated using dry mass ratio).
Table 5. Mission architecture matrix of Architecture-1.
Table 5. Mission architecture matrix of Architecture-1.
(BOS)No.RVD 1CEVMLOTSIMV 1 Δ v (m/s)
Launching to LEO1011100
LEO accelerating (leaving) (dv1)2011234011
Orbit Correction (OCC1) 2301123302
LMO accelerating (reaching) (dv2)4011232020
LMO stay5011100
Landing/Launching on Mars6001000
LMO docking 37111230.01
LMO accelerating (leaving) (dv3)8010232334
Orbit Correction (OCC2)901023117
1 RVD: Rendezvous and docking flag; IMV: Number of the active module. 2 OCC is considered about a 5% velocity increment of the entire transfer. 3 Orbit docking usually requires a 10 m/s velocity increment
Table 6. Mass scale of Architecture-1.
Table 6. Mass scale of Architecture-1.
ModuleCEVMLOTSTotalMax Mass
Original mass m 0 ( kg )60,00085,0002,207,8312,352,8312,207,831
Final mass m f ( kg )60,00085,000220,783365,783-
Table 7. Mission architecture matrix of Architecture-2.
Table 7. Mission architecture matrix of Architecture-2.
(BOS)No.RVDCEVMLOTS1OTS2IMV Δ v (m/s)
LEO10111100
dv1 (CEV)20102034011
dv1 (ML)30010244011
OCC1 (CEV)4010203302
OCC1 (ML)5001024302
dv2 (CEV)60102032020
dv2 (ML)70010242020
LMO stay80111100.01
Landing90010000
LMO docking101112030.01
dv3110102032334
OCC212010203117
Table 8. Mass scale of Architecture-2.
Table 8. Mass scale of Architecture-2.
ModuleCEVMLOTS1OTS2TotalMax Mass
m 0 ( kg )60,00085,0001,715,989502,4382,363,4271,715,989
m f ( kg )60,00085,000171,59950,244366,843-
Table 9. Mission architecture matrix of Architecture-3.
Table 9. Mission architecture matrix of Architecture-3.
(BOS)No.RVDCEVMLSMOTS1OTS2OTS3OTS4IMV Δ v (m/s)
LEO10111111100
dv1 (CEV)20100200044011
dv1 (ML)30010020054011
dv1(SM)40001002164011
OCC1 (CEV)5010020004302
OCC1 (ML)6001002005302
OCC1 (SM)7000100216302
dv2 (CEV)80100200042020
dv2 (ML)90010020052020
dv2 (SM)100001002162020
LMO stay110111111100
Landing120010000000
LMO docking131111000270.01
dv3140101000272334
OCC215010100027117
Table 10. Mass scale of Architecture-3.
Table 10. Mass scale of Architecture-3.
Module m 0 kg m f (kg)Module m 0 kg m f (kg)
CEV40,00040,000OTS3530,40953,041
ML85,00085,000OTS469,7326973
SM20,00020,000Total1,484,021278,902
OTS1236,44223,644Max mass530,409-
OTS2502,43850,244
Table 11. Mission architecture matrix of Architecture-4.
Table 11. Mission architecture matrix of Architecture-4.
(BOS)No.RVDCEVMLSMOTS1OTS2OTS3OTS4OTS5OTS6OTS7IMV Δ v (m/s)
LEO10111111111100
dv1 (CEV)20100210000044011
dv1 (ML)30010002100064011
dv1(SM)40001000021184011
OCC1 (CEV)5010002000005302
OCC1 (ML)6001000020007302
OCC1 (SM)7000100210219302
dv2 (CEV)80100020000052020
dv2 (ML)90010000200072020
dv2 (SM)100001000002192020
LMO stay110111010101100
Landing120010000000000
LMO docking1311110000002100.01
dv31401010000002102334
OCC2150101000000210117
Table 12. Mass scale of Architecture-4.
Table 12. Mass scale of Architecture-4.
Module m 0 (kg) m f (kg)Module m 0 (kg) m f (kg)
CEV40,00040,000OTS473,1817318
ML85,00085,000OTS5289,96028,996
SM20,00020,000OTS665,2386524
OTS1153,06415,306OTS755,7745577
OTS234,4383444Total1,141,917244,692
OTS3325,26132,526Max mass325,261-
Table 13. Mass scale of Architecture-5.
Table 13. Mass scale of Architecture-5.
ModuleCEVMLNTPTotalMax Mass
Original mass m 0 ( kg )60,00085,000250,788393,788250,788
Final mass m f ( kg )60,00085,00026,879171,879-
Table 14. Mass scale of Architecture-6.
Table 14. Mass scale of Architecture-6.
ModuleCEVMLNTP1NTP2TotalMax Mass
m 0 ( kg )60,00085,000147,139116,953409,092147,139
m f ( kg )60,00085,00016,51413,495175,009-
Table 15. Mass scale of Architecture-7.
Table 15. Mass scale of Architecture-7.
Module m 0 (kg) m f (kg)Module m 0 (kg) m f (kg)
CEV40,00040,000NTP369,6818768
ML85,00085,000NTP431,2234922
SM20,00020,000Total422,995179,999
NTP160,1377814Max mass116,953-
NTP2116,95313,495
Table 16. Mass scale of Architecture-8.
Table 16. Mass scale of Architecture-8.
Module m 0 (kg) m f (kg)Module m 0 (kg) m f (kg)
CEV40,00040,000NEP463,1086311
ML85,00085,000NTP549,6956770
SM20,00020,000NTP619,9463795
NTP142,4866049NTP727,8354584
NTP217,2343523Total418,905184,990
NTP383,89310,189Max mass85,000-
Table 17. Mission architecture matrix of Architecture-9.
Table 17. Mission architecture matrix of Architecture-9.
(BOS)No.RVDCEVMLNEPNTPIMV Δ v (m/s)
Launching to LEO10111100
LEO escape20111243175
LEO-LMO transfer30112136050
LMO entry40111241449
LMO stay50111100
Landing60010000
LMO docking71112130.01
LMO escape80101241449
LMO-Earth transfer90102135491
Table 18. Mass scale of Architecture-9.
Table 18. Mass scale of Architecture-9.
ModuleCEVMLNEPNTPTotalMax Mass
m 0 ( kg )60,00085,000166,648261,983573,631261,983
m f ( kg )60,00085,00060,00026,198231,198-
Table 19. Mission architecture matrix of Architecture-10.
Table 19. Mission architecture matrix of Architecture-10.
(BOS)No.RVDCEVMLSMNEPNTP1NTP2NTP3NTP4NTP5IMV Δ v (m/s)
LEO Launch1011111111100
LEO escape (M)2001020000043175
E-to-M (M)3001020000046050
LMO entry (M)4001020000041449
dv1 (C)5010002100054011
OCC1 (C)601000020006302
dv2 (C)7010000200062020
dv1 (S)8010000021174011
OCC1 (S)901000000218302
dv2 (S)10010000002182020
LMO stay11011110101100
Landing12001000000000
LMO docking13111110000290.01
dv3 (C)14010100000292334
OCC2 (C)1501010000029117
Table 20. Mass scale of Architecture-10.
Table 20. Mass scale of Architecture-10.
Module m 0 (kg) m f (kg)Module m 0 (kg) m f (kg)
CEV40,00040,000NTP349,6956770
ML85,00085,000NTP419,9463795
SM20,00020,000NTP527,8354584
NEP83,63160,000Total385,827229,721
NTP142,4866049Max mass85,000-
NTP217,2343523
Table 21. LEO Launch schedule.
Table 21. LEO Launch schedule.
LEO Launch No.Module (s)LEO Capacity Requirement (tons)
1ML85.0
2NTP383.9
3NTP1 + NTP259.7
4NTP463.1
5NTP5 + NTP669.6
6NTP7 + SM47.8
TotalAll modules418.9
Table 22. Detailed process of manned Mars exploration mission.
Table 22. Detailed process of manned Mars exploration mission.
TimeEvent
Before 5 September 2037All modules are launched to LEO and docked as 3 transfer combinations.
5 September 20373 combinations start Earth-to-Mars transfer, and NTP-1, 3, and 5 are abandoned after the first maneuver.
21 April 20383 combinations arrive at LMO, and NTP-2, 4, and 6 are abandoned after the second maneuver. All modules start preparations for Mars landing.
21 April 2038–18 July 2039CEV, ML, SM, and NTP-7 are docked together. The astronauts enter ML and then land on Mars, while the combination of CEV+SM+NTP-7 stays at LMO. After the landing mission, the astronauts go back to LMO using the upper stage and docking with CEV. Finally, the crew enters CEV and the upper stage is abandoned.
18 July 2039The combination of CEV+SM+NTP-7 starts Mars-to-Earth transfer.
15 May 2040The combination of CEV+SM+NTP-7 arrives at Earth. The astronauts enter the return cabin of CEV, re-enter the atmosphere and return to the ground.
Total mission time983 days
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shen, D.; Yue, Y.; Wang, X. Manned Mars Mission Analysis Using Mission Architecture Matrix Method. Aerospace 2022, 9, 604. https://doi.org/10.3390/aerospace9100604

AMA Style

Shen D, Yue Y, Wang X. Manned Mars Mission Analysis Using Mission Architecture Matrix Method. Aerospace. 2022; 9(10):604. https://doi.org/10.3390/aerospace9100604

Chicago/Turabian Style

Shen, Diyang, Yuxian Yue, and Xiaohui Wang. 2022. "Manned Mars Mission Analysis Using Mission Architecture Matrix Method" Aerospace 9, no. 10: 604. https://doi.org/10.3390/aerospace9100604

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