Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-24T15:00:14.945Z Has data issue: false hasContentIssue false

Effects of temperatures and high pressures on the growth and survivability of methanogens and stable carbon isotope fractionation: implications for deep subsurface life on Mars

Published online by Cambridge University Press:  02 August 2018

Navita Sinha
Affiliation:
Center for Space and Planetary Sciences, University of Arkansas, Fayetteville, AR72701, USA
Sudip Nepal
Affiliation:
Department of Physics, University of Arkansas, Fayetteville, AR72701, USA
Timothy Kral
Affiliation:
Center for Space and Planetary Sciences, University of Arkansas, Fayetteville, AR72701, USA Department of Biological Sciences, University of Arkansas, Fayetteville, AR72701, USA
Pradeep Kumar*
Affiliation:
Center for Space and Planetary Sciences, University of Arkansas, Fayetteville, AR72701, USA Department of Physics, University of Arkansas, Fayetteville, AR72701, USA
*
Author for correspondence: Pradeep Kumar, E-mail: pradeepk@uark.edu
Rights & Permissions [Opens in a new window]

Abstract

In order to examine the potential survivability of life in the Martian deep subsurface, we have investigated the effects of temperature (45°C, 55°C and 65°C) and pressure (1, 400, 800 and 1200 atm) on the growth, carbon isotopic data and morphology of chemolithoautotrophic anaerobic methanogenic archaea, Methanothermobacter wolfeii. The growth and survivability of this methanogen were determined by measuring the methane concentration in headspace gas samples after the cells were returned to their conventional growth conditions. Interestingly, this methanogen survived at all the temperatures and pressures tested. M. wolfeii demonstrated the highest methane concentration following exposure to pressure of 800 atm and a temperature of 65°C. We found that the stable carbon isotopic fractionation of methane, δ13C(CH4), was slightly more enriched in 12C at 1 atm and 55°C than the carbon isotopic data obtained in other temperature and pressure conditions. A comparison of the images of the cells before and after the exposure to different temperatures and pressures did not show any obvious alteration in the morphology of M. wolfeii. The research reported here suggests that at least one methanogen, M. wolfeii, may be able to survive under hypothetical Martian subsurface conditions with respect to temperature and pressure.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2018

Introduction

The current surface conditions of Mars are extremely harsh for any known life forms. Environmental factors such as low surface temperature, low atmospheric pressure, DNA damaging UV radiation and the presence of oxidizing compounds make Mars an inhospitable planet (Biemann et al., Reference Biemann, Oro, Toulmin, Orgel, Nier, Anderson, Simmonds, Flory, Diaz and Rushneck1977; Jakosky, Reference Jakosky1998; Cockell et al., Reference Cockell, Catling, Davis, Snook, Kepner, Lee and McKay2000). If life exists on the surface of Mars, it would have to challenge the environmental extremes of Mars and would have unique adaptation mechanisms. Therefore, the most feasible possibility of finding active known life forms on Mars would be ‘near or deep subsurface’, where the temperature and pressure will be higher compared with the surface and the environment will be protected from the damaging cosmic radiation. The discovery of extremophiles and the knowledge of Earth's subsurface biospheres have also bolstered the idea of searching for life in the subsurface of other planetary bodies such as Mars (Cavicchioli, Reference Cavicchioli2002).

The detection of methane in the Martian atmosphere (Formisano et al., Reference Formisano, Atreya, Encrenaz, Ignatiev and Giuranna2004; Krasnopolsky et al., Reference Krasnopolsky, Maillard and Owen2004; Mumma et al., Reference Mumma, Novak, DiSanti, Bonev and Dello Russo2004, Reference Mumma, Villanueva, Novak, Hewagama, Bonev, DiSanti, Mandell and Smith2009; Webster et al., Reference Webster, Mahaffy, Atreya, Flesch, Mischna, Meslin, Farley, Conrad, Christensen and Pavlov2015) has further reinforced the search for extinct or extant life on Mars. The reason is that most terrestrial methane is produced by biological sources either directly or indirectly (Atreya et al., Reference Atreya, Mahaffy and Wong2007). Methanogens are one of the various sources of methane on Earth. Some strains of methanogens have shown survivability and growth in simulated Martian physical and chemical conditions (Kral et al., Reference Kral, Bekkum and McKay2004, Reference Kral and Altheide2013, Reference Kral, Birch, Lavender and Virden2014, Reference Kral, Goodhart, Harpool, Hearnsberger, McCracken and McSpadden2015; McAllister and Kral, Reference McAllister and Kral2006; Sinha and Kral, Reference Sinha and Kral2015). For these reasons, methanogens, which are chemolithoautotrophic anaerobic archaea, have been considered ideal candidates for life on Mars (Boston et al., Reference Boston, Ivanov and McKay1992; Chapelle et al., Reference Chapelle, O'Neill, Bradley, Methé, Ciufo, Knobel and Lovley2002; Kral et al. Reference Kral, Bekkum and McKay2004, Reference Kral, Goodhart, Harpool, Hearnsberger, McCracken and McSpadden2015; Moran et al., Reference Moran, Miller, Kral and Scott2005; Chastain and Kral, Reference Chastain and Kral2012). The sources of methane on Mars are still unknown. Mars’ atmospheric methane could be the results of biotic, abiotic or a combination of both processes. Methanogens could be one of the several potential sources of Martian methane.

Stable carbon isotope fractionation is one of the several potential techniques to differentiate between biogenic and abiogenic sources of methane (Allen et al., Reference Allen, Sherwood Lollar, Runnegar, Oehler, Lyons, Manning and Summers2006). Stable carbon isotope fractionation data for terrestrial atmospheric methane have been used in order to understand the environments, pathways and origins or substrates of methanogenesis (Rothschild and DesMarais, Reference Rothschild and DesMarais1989; Schidlowski, Reference Schidlowski1992; Londry et al., Reference Londry, Dawson, Grover, Summons and Bradley2008). Sinha and Kral have recently studied the carbon isotope fractionation following methanogenesis on various Mars regolith analogues and found enriched values of 12C on the clay called montmorillonite compared with the carbon isotopic data obtained on other Mars analogues such as JSC Mars-1, Artificial Mars Simulant and Mojave Mars Simulant (Sinha and Kral, Reference Sinha and Kral2015).

Analogous to Earth's subsurface environments, hydrothermal systems might have existed and may also be present on Mars. On Earth, several hyperthermophilic and barophilic archaea have been isolated and characterized from the deep sea floor and hydrothermal sites (Canganella et al., Reference Canganella, Gonzalez, Yanagibayashi, Kato and Horikoshi1997; Horikoshi, Reference Horikoshi1998; Takai et al., Reference Takai, Inoue and Horikoshi2002; Shimizu et al., Reference Shimizu, Upadhye, Ishijima and Naganuma2011). Some of the most abundant species near hydrothermal vents are hyperthermophilic methanogens (Takai et al., Reference Takai, Gamo, Tsunogai, Nakayama, Hirayama, Nealson and Horikoshi2004). Life near a hydrothermal vent experiences a wide range of temperature and pressure. It has been found that a methanogen of genus Methanopyrus can grow up to a temperature of 110°C (Takai et al., Reference Takai, Nakamura, Toki, Tsunogai, Miyazaki, Miyazaki, Hirayama, Nakagawa, Nunoura and Horikoshi2008) and Methanococcus jannaschii demonstrated methanogenesis up to a pressure of 750 atm (Miller et al., Reference Miller, Shah, Nelson, Ludlow and Clark1988). Methanogenesis has also been found in organisms thriving in 3.5-million-year-old subseafloor basalt on Earth and was detected by using δ13C data (Lever et al., Reference Lever, Rouxel, Alt, Shimizu, Ono, Coggon, Shanks, Lapham, Elvert and Prieto-Mollar2013). Several surface features, geochemical and isotopic evidence in a Martian meteorite point to the activity of hydrothermal systems on Mars (Brakenridge et al., Reference Brakenridge, Newsom and Baker1985; Romanek et al., Reference Romanek, Grady, Wright, Mittlefehldt, Socki, Pillinger and Gibson1994; Shock, Reference Shock1997; Watson et al., Reference Watson, Hutcheon, Epstein and Stolper1994). Therefore, similar subsurface biota might exist in the Mars’ subsurface.

There are a few studies done on the effect of temperature and pressure on various strains of methanogens. However, the effects of a wide range of temperature and high pressure on methanogens in the context of Mars’ subsurface have not been studied before. The goal of this study is to examine the growth and survivability of a Mars’ model organism, Methanothermobacter wolfeii in a wide range of temperatures (45–65°C) and pressures (1–1200 atm). M. wolfeii, a hydrogenotrophic methanogenic archaeon used in this study, utilizes CO2 for its carbon source, H2 for its energy source and produces methane as a metabolic byproduct. We have also measured the stable carbon isotope fractionation of methane, δ13C(CH4), in order to understand the effect of optimal and non-optimal temperature and pressure on the carbon isotopic data. The images of cells were also acquired and were analysed to investigate morphological changes following exposure to various pressures and temperatures.

Materials and methods

Preparation of a stock culture of M. wolfeii

M. wolfeii (OCM36) was obtained from the Oregon Collection of Methanogens, Portland State University, Portland, OR, USA and was grown in MM medium consisting of potassium phosphate, ammonium chloride, calcium chloride, resazurin as an oxygen indicator and many trace minerals (Xun et al., Reference Xun, Boone and Mah1988) in a bicarbonate buffer. MM medium was prepared in an anaerobic chamber (Coy Laboratory Products Inc., Grass Lake Charter Township, MI, USA), which was filled with 90% carbon dioxide and 10% hydrogen. The growth medium was then transferred into anaerobic culture tubes inside the anaerobic chamber as described previously (Boone et al., Reference Boone, Johnson and Liu1989). The tubes were then sealed with butyl rubber stoppers, removed from the chamber, crimped with aluminium caps and autoclaved for sterilization.

A sterile sodium sulphide solution (2.5% wt/vol; 0.15 ml per 10 ml media) was then added to each tube about an hour prior to inoculation of the methanogen (Boone et al. Reference Boone, Johnson and Liu1989) in order to eliminate any residual molecular oxygen from the tubes containing the media. After inoculating the methanogen, the tubes were pressurized with 2 atm of H2 gas and incubated at the organism's optimal growth temperature, 55°C. The stock culture of M. wolfeii was maintained by transferring into fresh MM medium every 15 days.

The growth and survivability of this methanogen were determined by measuring the methane concentration in an aliquot of the headspace gas using a gas chromatograph (Shimadzu 2014, Lenexa, KS, USA), which was equipped with a flame ionization detector.

Temperature–pressure experiments

The schematic diagram of the temperature–pressure chamber is shown in Fig. 1. A quartz cuvette (Spectracell, Oreland, PA, USA) was filled with 1 ml of fresh liquid culture of M. wolfeii and was capped with a Teflon cap (E. I. DuPont de Nemours, Paris, France) inside the anaerobic chamber. The cuvette filled with sample was then put into a high hydrostatic pressure–temperature chamber (ISS, Champaign, IL, USA), which was filled with water and was connected to a piston. The pressure inside the chamber was developed with the help of a piston by pressurizing liquid water, and the pressure was measured with a pressure gauge attached to the piston (Kumar and Libchaber, Reference Kumar and Libchaber2013). The temperature of the chamber was maintained using a circulating water bath (Neslab, USA) and was measured in real time using a thermocouple (National Instruments, USA) attached to the chamber. Before loading the sample, the temperature of the chamber was equilibrated to the desired temperature and the pressure was applied after the sample cuvette was put into the high-temperature–pressure chamber. The system equilibrated to the desired pressure and temperature in <2 min after loading the sample. In this work, we preformed experiments at three different temperatures − 45°C, 55°C and 65°C and four different pressures – 1, 400, 800 and 1200 atm; resulting in a total of 12 sets of experiments. For a given temperature and pressure, the sample was kept in the temperature–pressure chamber for 15 h. After this time, the pressure of the chamber was released and the sample-filled cuvette was removed from the chamber.

Fig. 1. Experimental setup for the temperature–pressure experiment.

The cells from the cuvette were then transferred to fresh medium. Five hundred microlitres of cells from the cuvette were mixed with 500 µl of sterilized MM medium in a vial to make a total of 1 ml of culture. From this, 300 µl cells were inoculated into three different anaerobic tubes containing 10 ml of sterilized MM medium to make triplicate samples. The tubes were then pressurized with 2 atm H2 gas and incubated at the conventional growth temperature, 55°C. The growth and survivability were determined by measuring methane concentration in the headspace gas of each sample at regular intervals with the help of a gas chromatograph (Shimadzu 2014).

Determination of stable carbon isotope fractionation

Using the procedure described by Sinha and Kral (Reference Sinha and Kral2015), the carbon isotope fractionation of methane in the headspace gas of all samples was measured periodically by a Piccaro Cavity Ringdown Spectrometer G-2201-I isotopic CO2/CH4 in the University of Arkansas Isotope laboratory. The carbon isotope fractionation, δ13C, was calculated using the following equation:

(1)$${\rm \delta} ^{13}{\rm C}_{{\rm sample}} = \left\{ {\displaystyle{{{ \left({13_{\rm c} \over 12_{\rm c}}\right)}_{\rm sample}} \over {{ \left({ 13_{\rm c} \over 12_{\rm c} }\right) }_{{\rm reference}}}} - 1} \right\} \times 1000.$$

The reference isotopic standard for δ13C is Pee Dee Belemonite (O'Leary, Reference O'Leary1981).

Imaging of cells

Phase-contrast images of the cells were obtained before and after exposure to various temperatures and pressures using a SPOT Imaging camera and 40× objective mounted on a Nikon Optiphot microscope.

Results

M. wolfeii was exposed to temperatures of 45°C, 55°C and 65°C and pressures of 1, 400, 800 and 1200 atm in a high hydrostatic pressure–temperature chamber for 15 h. Interestingly, M. wolfeii survived at all temperatures and pressures studied here. This methanogenic archaeon demonstrated methanogenesis by producing methane after the cells were returned to their conventional growth conditions. All measurements of methane concentration were taken post exposure to different temperatures and pressures.

In Fig. 2, we show methane concentration, [CH4], as a function of time for the cells exposed to the temperatures − 45°C, 55°C and 65°C. Data in these figures represent the average methane concentration produced by methanogens exposed to the pressures – 1, 400, 800 and 1200 atm. For all temperatures and pressures, samples were in triplicates. For each temperature, M. wolfeii demonstrated highest methane concentration following the exposure to 800 atm of pressure and the lowest methane concentration following 1 atm. Highest methane concentration varied for different temperatures and reaches the maximum on different days. For example, at 800 atm and 45°C and 55°C, methane concentration reached its maximum on the third day, while for 65°C, the methane concentration reached its maximum on the fourth day.

Fig. 2. Methane concentration as a function of time following methanogenesis for Methanothermobacter wolfeii exposed to the temperatures − 45°C, 55°C and 65°C. The different coloured lines in these figures represent the average methane concentration produced by methanogens exposed to the pressures – 1, 400, 800 and 1200 atm.

Moreover, we found that at 55°C, the optimal growth temperature of M. wolfeii, the lag phase was little more than 24 h, whereas at 45°C and 65°C (non-optimal growth temperatures), the durations of lag phase were <24 h. This suggests that after the release of high-pressure stress, M. wolfeii adapted quickly to the optimal pressure condition at the non-optimal growth temperatures as compared with the optimal growth temperature. One plausible reason for different duration of lag phase observed could be due to the kinetic of reversibility of cells after they are brought back to favourable pressure after being exposed to high pressure. A recent study on Escherichia coli (Nepal and Kumar, Reference Nepal and Kumar2018) suggests that the rate at which these cells revert back to normal pressure phenotype after being exposed to high pressure depends on the magnitude and the duration of applied pressure. It is likely that the rate of reversibility to normal pressure phenotype is non-monotonic with pressure.

In Fig. 3, we show methane concentration, [CH4], as a function of time for the cells exposed to the pressures – 1, 400, 800 and 1200 atm. Data in these figures represent methane produced by methanogens exposed to the temperatures − 45°C, 55°C and 65°C. For all temperatures and pressures, samples were in triplicates. For each pressure, M. wolfeii exhibited the highest methane concentration after exposure to 65°C. For the lowest pressure, 1 atm, the methane concentration reached its maximum on the third day following a temperature of 65°C. However, for the highest pressure, 1200 atm, the methane concentration reached its maximum on the fourth day for all of the temperatures tested. The decay observed after 4–5 days arises due to saturation of growth of the cells due to limited amount of nutrients. Since the measurement of methane in the sample requires us to take out a constant small volume of the headspace gas, it results in decay of methane concentration after saturation since no further methane is being produced by the cells. If the growth is faster after being exposed to a given pressure, the saturation will reach faster and hence the decay will be observed earlier, which can be seen in Figs. 2 and 3.

Fig. 3. Methane concentration as a function of time following methanogenesis for Methanothermobacter wolfeii exposed to the pressures – 1, 400, 800 and 1200 atm. The different coloured lines in these figures represent the average methane concentration produced by methanogens exposed to the temperatures − 45°C, 55°C and 65°C.

We next studied the effect of temperature and pressure on the stable carbon isotope fractionation of methane produced by M. wolfeii. In Table 1, we have listed the carbon isotope fractionation data obtained on the second and sixth day following the pressure–temperature exposures. For 1 atm and 55°C (conventional growth conditions of M. wolfeii), δ13C(CH4) on the second and sixth day are −74.36‰ and −73.35‰, which are comparable with the published data for standard growth conditions (Sinha and Kral, Reference Sinha and Kral2015). Here we found slightly depleted values of δ13C(CH4) in the conventional conditions compared with the δ13C(CH4) data obtained at non-optimal temperatures and elevated pressures.

Table 1. Carbon isotope fractionation of methane, δ13C(CH4), produced by Methanothermobacter wolfeii obtained on day 2 and day 6 following different temperature–pressure exposures

Values shown are in per mil.

In order to compare any morphological changes in the cells exposed to optimal and non-optimal pressures and temperature, we obtained phase-contrast images of the cells before and after exposure to different temperatures and pressures. In Fig. 4, we compare the images of cells exposed to temperatures − 55°C and 65°C and pressures – 1and 800 atm. We did not find any significant effect of temperature and pressure on the morphology of M. wolfeii. It has been shown that high pressure exhibits cell division inhibiting the effects on mesophilic bacteria such as E. coli (Kumar and Libchaber, Reference Kumar and Libchaber2013). The lack of morphological alteration such as elongation in the case of M. wolfeii suggests that the exposure time of 15 h was not long enough for the cells to undergo multiple cycles of cell division.

Fig. 4. Images of Methanothermobacter wolfeii after exposure to (a) T = 55°C and P = 1 atm and (b) T = 65°C and P = 1200 atm under a magnification of 400×.

Discussion

The growth and survivability of several methanogenic archaea in simulated Martian surface conditions such as at low temperature, low pressure and desiccation have been studied previously (Kral et al., Reference Kral, Brink, Miller and McKay1998; Kendrick and Kral, Reference Kendrick and Kral2006; Reid et al., Reference Reid, Sparks, Lubow, McGrath, Livio, Valenti, Sowers, Shukla, MacAuley and Miller2006; Kral et al., Reference Kral, Altheide, Lueders and Schuerger2011). On the other hand, deep subsurface of Mars could potentially offer a feasible environment for a biosphere. The challenges are then high pressure, temperature and availability of liquid water, nutrients and the source of energy. It is imperative to locate liquid water in the subsurface of Mars. All terrestrial life needs water at some stage in their life cycle. Martian geophysical models suggest that the liquid water in the subsurface of Mars could be present to a depth of ~310 km (Jones et al., Reference Jones, Lineweaver and Clarke2011). According to this model, the depth and pressure for approximate temperature range 45–65°C would be between 1–30 km and 100–3000 atm.

In this work, we have investigated the effects of temperature (45°C, 55°C and 65°C) and pressure (1, 400, 800 and 1200 atm) on the growth, carbon isotopic data and morphology of M. wolfeii. The growth and survivability of methanogens were determined by measuring methane concentration in the headspace gas samples after they were returned to their conventional growth conditions. Due to the limitation on the maximum temperature and prolonged experiments in hydrostatic temperature–pressure chamber, 15 h long experiments were performed with a maximum temperature of 65°C and pressure of 1200 atm. Since the typical doubling time for M. Wolfeii is about 10 h at conventional growth conditions, our experiments should be able to detect the survivability over the timescale of our experiments. We found that M. wolfeii was able to endure a temperature of 65°C and a pressure of 1200 atm for at least that amount of time. Surprisingly, M. wolfeii demonstrated methanogenesis following exposure to all the temperatures and pressures studied here. M. wolfeii exhibited the highest methane concentration following exposure to a pressure of 800 tm and a temperature of 65°C. The exact reason for this is not clear and will be the focus of future studies.

The stable carbon isotope fractionation of methane was measured in different temperature and pressure experiments. We found very slight or negligible difference in the carbon isotopic data following optimal and non-optimal growth conditions of M. wolfeii. The δ13C(CH4) data in the optimal conditions were slightly lower than the δ13C(CH4) data in non-optimal conditions. It is possible that for M. wolfeii, 15 h exposure to various temperatures and pressures may not be long enough to have a significant effect on the carbon isotopic data. Takai et al. (Reference Takai, Nakamura, Toki, Tsunogai, Miyazaki, Miyazaki, Hirayama, Nakagawa, Nunoura and Horikoshi2008) have found that Methanopyrus kandleri produced isotopically heavier methane under high hydrostatic pressure conditions compared with the methane produced by M. kandleri in a conventional growth condition (Takai et al., Reference Takai, Nakamura, Toki, Tsunogai, Miyazaki, Miyazaki, Hirayama, Nakagawa, Nunoura and Horikoshi2008).

We compared the morphology of cells before and after the exposure to different temperatures and pressures. Most of the cells remained intact and we did not find obvious alteration in the morphology of M. wolfeii, unlike E. coli, which shows increase in cell length with increase in pressure (Kumar and Libchaber, Reference Kumar and Libchaber2013). In Figs. 5(a) and (b), we show the probability distribution, ρ(l), of cell length, l, at P = 1 atm and P = 1200 atm. We find the mean length <l> of the cells to be 4.15 ± 0.32 μm at P = 1 atm and 3.98 ± 0.30 μm at P = 1200 atm. The error bars on the mean length of the cells is estimated as the sum of errors arising from segmentation of the images and standard error due to finite sampling. The number of cells used to obtain the distribution were 597 and 670 for P = 1 atm and 1200 atm, respectively. The variance ${\rm \sigma} _l^2 $ and the coefficient of variation, defined as (σ/〈l〉), for the distributions are 2.30 and 0.36 at P = 1 atm and 1.76 and 0.33 at P = 1200 atm, respectively. Our results suggest that both the distribution and the mean cell length does not change appreciably with pressure. Since the doubling time of E. coli is very small as compared with M. wolfeii, it was possible to detect stochasticity in cell division and elongation of E. coli at high pressure over a large number of generation times.

Fig. 5. Probability distributions, ρ(l), of cell length, l, for pressure (a) P=1 atm and (b) P=1200 atm.

The results presented here suggest that one Mars’ model microorganism, M. wolfeii, can survive under presumed Martian subsurface conditions in terms of temperature and pressure. Therefore, the search for life on Mars should also be focused on the deep subsurface of Mars.

Conclusions

Methanogens have been considered ideal life forms on Mars for a long time. Here, we have examined the growth and the survivability of a methanogen, M. wolfeii, in presumed deep-subsurface environments in terms of temperature and pressure. We used three different temperatures (45, 55 and 65°C) and four different pressures (1, 400, 800 and 1200 atm). M. wolfeii demonstrated survivability by producing methane following exposure to all different temperatures and pressures. The growth and survivability of M. wolfeii were investigated after returning them to their conventional growth conditions. We have also measured the carbon isotopic fractionation of methane produced by M. wolfeii and found that δ13C(CH4) in optimal growth conditions was slightly lower than the values obtained in non-optimal growth conditions. A comparison of the images of cells before and after the exposure to different temperatures and pressures did not reveal any apparent alteration in the morphology of M. wolfeii.

Acknowledgements

The authors would like to thank Erik Pollock, manager of the University of Arkansas Stable Isotope Laboratory, for his assistance in obtaining isotope data.

Conflict of interest

None.

References

Allen, M, Sherwood Lollar, B, Runnegar, B, Oehler, DZ, Lyons, JR, Manning, CE and Summers, ME (2006) Is Mars alive? Eos. Transactions American Geophysical Union 87, 433439.CrossRefGoogle Scholar
Atreya, SK, Mahaffy, PR and Wong, A-S (2007) Methane and related trace species on Mars: origin, loss, implications for life, and habitability. Planetary and Space Science 55, 358369.CrossRefGoogle Scholar
Biemann, K, Oro, J, Toulmin, P, Orgel, L, Nier, A, Anderson, D, Simmonds, P, Flory, D, Diaz, A and Rushneck, D (1977) The search for organic substances and inorganic volatile compounds in the surface of Mars. Journal of Geophysical Research 82, 46414658.CrossRefGoogle Scholar
Boone, DR, Johnson, RL and Liu, Y (1989) Diffusion of the interspecies electron carriers H2 and formate in methanogenic ecosystems and its implications in the measurement of Km for H2 or formate uptake. Applied and Environmental Microbiology 55, 17351741.CrossRefGoogle ScholarPubMed
Boston, PJ, Ivanov, MV and McKay, CP (1992) On the possibility of chemosynthetic ecosystems in subsurface habitats on Mars. Icarus 95, 300308.CrossRefGoogle ScholarPubMed
Brakenridge, GR, Newsom, HE and Baker, VR (1985) Ancient hot springs on Mars: origins and paleoenvironmental significance of small Martian valleys. Geology 13, 859862.2.0.CO;2>CrossRefGoogle Scholar
Canganella, F, Gonzalez, JM, Yanagibayashi, M, Kato, C and Horikoshi, K. (1997) Pressure and temperature effects on growth and viability of the hyperthermophilic archaeon Thermococcus peptonophilus. Archives of Microbiology 168, 17.CrossRefGoogle ScholarPubMed
Cavicchioli, R (2002) Extremophiles and the search for extraterrestrial life. Astrobiology 2, 281292.CrossRefGoogle ScholarPubMed
Chapelle, FH, O'Neill, K, Bradley, PM, Methé, BA, Ciufo, SA, Knobel, LL and Lovley, DR (2002) A hydrogen-based subsurface microbial community dominated by methanogens. Nature 415, 312315.CrossRefGoogle ScholarPubMed
Chastain, B and Kral, T (2012) Methane production by methanogens in the presence of Mars regolith analogs (abstract 271) in American Society for Microbiology General Meeting. San Francisco.Google Scholar
Cockell, CS, Catling, DC, Davis, WL, Snook, K, Kepner, RL, Lee, P and McKay, CP (2000) The ultraviolet environment of Mars: biological implications past, present, and future. Icarus 146, 343359.CrossRefGoogle ScholarPubMed
Formisano, V, Atreya, S, Encrenaz, T, Ignatiev, N and Giuranna, M (2004) Detection of methane in the atmosphere of Mars. Science 306, 17581761.CrossRefGoogle ScholarPubMed
Horikoshi, K (1998) Barophiles: deep-sea microorganisms adapted to an extreme environment. Current Opinion in Microbiology 1, 291295.CrossRefGoogle Scholar
Jakosky, B (1998) The Search for Life on Other Planets. Cambridge, UK: Cambridge University Press.Google Scholar
Jones, EG, Lineweaver, CH and Clarke, JD (2011) An extensive phase space for the potential Martian biosphere. Astrobiology 11, 10171033.CrossRefGoogle ScholarPubMed
Kendrick, MG and Kral, TA (2006) Survival of methanogens during desiccation: implications for life on Mars. Astrobiology 6, 546551.CrossRefGoogle ScholarPubMed
Kral, TA and Altheide, ST (2013) Methanogen survival following exposure to desiccation, low pressure and Martian regolith analogs. Planetary and Space Science 89, 167171.CrossRefGoogle Scholar
Kral, TA, Brink, KM, Miller, SL and McKay, CP (1998) Hydrogen consumption by methanogens on the early Earth. Origins of Life and Evolution of the Biosphere 28, 311319.CrossRefGoogle ScholarPubMed
Kral, TA, Bekkum, CR and McKay, CP (2004) Growth of methanogens on a Mars soil simulant. Origins of Life and Evolution of the Biosphere 34, 615626.CrossRefGoogle ScholarPubMed
Kral, TA, Altheide, TS, Lueders, AE and Schuerger, AC. (2011) Low pressure and desiccation effects on methanogens: implications for life on Mars. Planetary and Space Science 59, 264270.CrossRefGoogle Scholar
Kral, TA, Birch, W, Lavender, LE and Virden, BT (2014) Potential use of highly insoluble carbonates as carbon sources by methanogens in the subsurface of Mars. Planetary and Space Science 101, 181185.CrossRefGoogle Scholar
Kral, TA, Goodhart, TH, Harpool, JD, Hearnsberger, CE, McCracken, GL and McSpadden, SW (2015) Sensitivity and adaptability of methanogens to perchlorates: implications for life on Mars. Planetary and Space Science 120, 8795.CrossRefGoogle Scholar
Krasnopolsky, VA, Maillard, JP and Owen, TC (2004) Detection of methane in the Martian atmosphere: evidence for life? Icarus 172, 537547.CrossRefGoogle Scholar
Kumar, P and Libchaber, A (2013) Pressure and temperature dependence of growth and morphology of Escherichia coli: experiments and stochastic model. Biophysical Journal 105, 783793.CrossRefGoogle ScholarPubMed
Lever, MA, Rouxel, O, Alt, JC, Shimizu, N, Ono, S, Coggon, RM, Shanks, WC, Lapham, L, Elvert, M and Prieto-Mollar, X (2013) Evidence for microbial carbon and sulfur cycling in deeply buried ridge flank basalt. Science 339, 13051308.CrossRefGoogle ScholarPubMed
Londry, KL, Dawson, KG, Grover, HD, Summons, RE and Bradley, AS (2008) Stable carbon isotope fractionation between substrates and products of Methanosarcina barkeri. Organic Geochemistry 39, 608621.CrossRefGoogle Scholar
McAllister, SA and Kral, TA (2006) Methane production by methanogens following an aerobic washing procedure: simplifying methods for manipulation. Astrobiology 6, 819823.CrossRefGoogle ScholarPubMed
Miller, JF, Shah, NN, Nelson, CM, Ludlow, JM and Clark, DS. (1988) Pressure and temperature effects on growth and methane production of the extreme thermophile Methanococcus jannaschii. Applied and Environmental Microbiology 54, 30393042.CrossRefGoogle ScholarPubMed
Moran, M, Miller, JD, Kral, T and Scott, D (2005) Desert methane: implications for life detection on Mars. Icarus 178, 277280.CrossRefGoogle Scholar
Mumma, M, Novak, R, DiSanti, M, Bonev, B and Dello Russo, N. (2004) Detection and mapping of methane and water on Mars. Bulletin of the American Astronomical Society 36, 1127.Google Scholar
Mumma, MJ, Villanueva, GL, Novak, RE, Hewagama, T, Bonev, BP, DiSanti, MA, Mandell, AM and Smith, MD (2009) Strong release of methane on Mars in northern summer 2003. Science 323, 10411045.CrossRefGoogle ScholarPubMed
Nepal, S and Kumar, P (2018) Dynamics of phenotypic switching of bacterial cells with temporal fluctuations in pressure. Physical Review E 97, 052411.CrossRefGoogle ScholarPubMed
O'Leary, MH (1981) Carbon isotope fractionation in plants. Phytochemistry 20, 553567.CrossRefGoogle Scholar
Reid, I, Sparks, W, Lubow, S, McGrath, M, Livio, M, Valenti, J, Sowers, K, Shukla, H, MacAuley, S and Miller, T (2006) Terrestrial models for extraterrestrial life: methanogens and halophiles at Martian temperatures. International Journal of Astrobiology 5, 8997.CrossRefGoogle Scholar
Romanek, CS, Grady, MM, Wright, I, Mittlefehldt, D, Socki, R, Pillinger, C and Gibson, E (1994) Record of fluid rock interactions on Mars from the meteorite ALH84001. Nature 372, 655657.CrossRefGoogle ScholarPubMed
Rothschild, L and DesMarais, D (1989) Stable carbon isotope fractionation in the search for life on early Mars. Advances in Space Research 9, 159165.CrossRefGoogle ScholarPubMed
Schidlowski, M (1992) Stable carbon isotopes: possible clues to early life on Mars. Advances in Space Research 12, 101110.CrossRefGoogle ScholarPubMed
Shimizu, S, Upadhye, R, Ishijima, Y and Naganuma, T (2011) Methanosarcina horonobensis sp. nov., a methanogenic archaeon isolated from a deep subsurface Miocene formation. International Journal of Systematic and Evolutionary Microbiology 61, 25032507.CrossRefGoogle ScholarPubMed
Shock, EL (1997) High-temperature life without photosynthesis as a model for Mars. Journal of Geophysical Research 102, 2368723694.CrossRefGoogle ScholarPubMed
Sinha, N and Kral, TA (2015) Stable carbon isotope fractionation by methanogens growing on different Mars regolith analogs. Planetary and Space Science 112, 3541.CrossRefGoogle Scholar
Takai, K, Inoue, A and Horikoshi, K (2002) Methanothermococcus okinawensis sp. nov., a thermophilic, methane-producing archaeon isolated from a Western Pacific deep-sea hydrothermal vent system. International Journal of Systematic and Evolutionary Microbiology 52, 10891095.Google ScholarPubMed
Takai, K, Gamo, T, Tsunogai, U, Nakayama, N, Hirayama, H, Nealson, KH and Horikoshi, K (2004) Geochemical and microbiological evidence for a hydrogen-based, hyperthermophilic subsurface lithoautotrophic microbial ecosystem (HyperSLiME) beneath an active deep-sea hydrothermal field. Extremophiles 8, 269282.CrossRefGoogle ScholarPubMed
Takai, K, Nakamura, K, Toki, T, Tsunogai, U, Miyazaki, M, Miyazaki, J, Hirayama, H, Nakagawa, S, Nunoura, T and Horikoshi, K (2008) Cell proliferation at 122 C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. Proceedings of the National Academy of Sciences 105, 1094910954.CrossRefGoogle Scholar
Watson, LL, Hutcheon, ID, Epstein, S and Stolper, EM (1994) Water on Mars: clues from deuterium/hydrogen and water contents of hydrous phases in SNC meteorites. Science 265, 8690.CrossRefGoogle ScholarPubMed
Webster, CR, Mahaffy, PR, Atreya, SK, Flesch, GJ, Mischna, MA, Meslin, P-Y, Farley, KA, Conrad, PG, Christensen, LE and Pavlov, AA (2015) Mars methane detection and variability at Gale crater. Science 347, 415417.CrossRefGoogle ScholarPubMed
Xun, L, Boone, DR and Mah, RA (1988) Control of the life cycle of Methanosarcina mazei S-6 by manipulation of growth conditions. Applied and Environmental Microbiology 54, 20642068.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Experimental setup for the temperature–pressure experiment.

Figure 1

Fig. 2. Methane concentration as a function of time following methanogenesis for Methanothermobacter wolfeii exposed to the temperatures − 45°C, 55°C and 65°C. The different coloured lines in these figures represent the average methane concentration produced by methanogens exposed to the pressures – 1, 400, 800 and 1200 atm.

Figure 2

Fig. 3. Methane concentration as a function of time following methanogenesis for Methanothermobacter wolfeii exposed to the pressures – 1, 400, 800 and 1200 atm. The different coloured lines in these figures represent the average methane concentration produced by methanogens exposed to the temperatures − 45°C, 55°C and 65°C.

Figure 3

Table 1. Carbon isotope fractionation of methane, δ13C(CH4), produced by Methanothermobacter wolfeii obtained on day 2 and day 6 following different temperature–pressure exposures

Figure 4

Fig. 4. Images of Methanothermobacter wolfeii after exposure to (a) T = 55°C and P = 1 atm and (b) T = 65°C and P = 1200 atm under a magnification of 400×.

Figure 5

Fig. 5. Probability distributions, ρ(l), of cell length, l, for pressure (a) P=1 atm and (b) P=1200 atm.