Hostname: page-component-745bb68f8f-cphqk Total loading time: 0 Render date: 2025-02-06T09:56:28.408Z Has data issue: false hasContentIssue false

An active nitrogen cycle on Mars sufficient to support a subsurface biosphere

Published online by Cambridge University Press:  16 January 2012

C.S. Boxe*
Affiliation:
Earth and Space Science Division, Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA
K.P. Hand
Affiliation:
Earth and Space Science Division, Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA
K.H. Nealson
Affiliation:
Department of Earth Sciences, University of Southern California, Los Angeles, CA 90089, USA
Y.L. Yung
Affiliation:
Division of Geological and Planetary Sciences, California Institute of Technology, 1200 East California Boulevard, Pasadena, CA 91125, USA
A. Saiz-Lopez
Affiliation:
Earth and Space Science Division, Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA Laboratory for Atmospheric and Climate Sciences, CSIC, Toledo, Spain
Rights & Permissions [Opens in a new window]

Abstract

Mars' total atmospheric nitrogen content is 0.2 mbar. One-dimensional (1D) photochemical simulations of Mars' atmosphere show that nitric acid (HNO3(g)), the most soluble nitrogen oxide, is the principal reservoir species for nitrogen in its lower atmosphere, which amounts to a steady-state value of 6×10−2 kg or 4 moles, conditions of severe nitrogen deficiency. Mars could, however, support ∼1015 kg of biomass (∼1 kg N m−2) from its current atmospheric nitrogen inventory. The terrestrial mass ratio of nitrogen in biomass to that in the atmosphere is ∼10−5; applying this ratio to Mars yields ∼1010 kg of total biomass – also, conditions of severe nitrogen deficiency. These amounts, however, are lower limits as the maximum surface-sink of atmospheric nitrogen is 2.8 mbar (9×1015 kg of N), which indicates, in contradistinction to the Klingler et al. (1989), that biological metabolism would not be inhibited in the subsurface of Mars. Within this context, we explore HNO3 deposition on Mars' surface (i.e. soil and ice-covered regions) on pure water metastable thin liquid films. We show for the first time that the negative change in Gibbs free energy increases with decreasing HNO3(g) (NO3(aq)) in metastable thin liquid films that may exist on Mars' surface. We also show that additional reaction pathways are exergonic and may proceed spontaneously, thus providing an ample source of energy for nitrogen fixation on Mars. Lastly, we explore the dissociation of HNO3(g) to form NO3(aq) in metastable thin liquid films on the Martian surface via condensed phase simulations. These simulations show that photochemically produced fixed nitrogen species are not only released from the Martian surface to the gas-phase, but more importantly, transported to lower depths from the Martian surface in transient thin liquid films. A putative biotic layer at 10 m depth would produce HNO3 and N2 sinks of −54 and −5×1012 molecules cm−2 s−1, respectively, which is an ample supply of available nitrogen that can be efficiently transported to the subsurface. The downward transport as well as the release to the atmosphere of photochemically produced fixed nitrogen species (e.g. NO2, NO and NO2) suggests the existence of a transient but active nitrogen cycle on Mars.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2012

Introduction

Dinitrogen (N2) comprises 2.7% of the Martian atmosphere, which is equivalent to a total atmospheric mass of ∼1015 kg. Still, to date, there has been no direct and thorough analysis of the nitrogen content of the Martian lithosphere. Consequently, the total inventory of nitrogen for the planet is unknown (Mancinelli & Banin Reference Mancinelli and Banin2003; Summers & Khare Reference Summers and Khare2007), which contributes to the degree of uncertainty pertaining to the amount of nitrogen lost to space compared to the amount deposited on its surface and sequestered at its subsurface. It may be possible, though, that both processes have taken place concurrently and are ongoing (Mancinelli & Banin Reference Mancinelli and Banin2003; Summers & Khare Reference Summers and Khare2007). Moreover, the isotope ratio enhancement of atmospheric nitrogen on Mars compared to Earth (15N/14N (Mars)/15N/14N (Earth) ∼1.6) is used as evidence for significant nitrogen loss but not the total amount lost to space (Jakosky & Phillips Reference Jakosky and Phillips2001).

Thin liquid films on the surface and subsurface (i.e. tens of centimetres to 1 m depth) of Mars may range from 0.2 to 5 nm and ∼1 nm, respectively (Boxe et al. submitted). Although the diameter of bacteria, archaea and eukarya range from ∼0.40 μm to several microns (Huber et al. Reference Huber2002; Price Reference Price2007), biological activity on Martian surface or subsurface may not be limited by the nanometre dimensions of such thin liquid films (Price Reference Price2007). As nitrogen is an essential biochemical element, we discuss the implications for nitrogen fixation on Mars due to the existence of such films. We calculate upper and lower limits for Martian steady-state biomass by assuming the abundance of gas-phase nitrogen to be the limiting factor for biomass production. We also consider a nitrogen cycle facilitated by nitric acid (HNO3(g)) dissolution and subsequent heterogeneous photochemistry in thin films of metastable liquid water that may exist on the surface and subsurface of Mars; the heterogeneous photochemistry of nitrate (NO3) produces NO, NO2 and NO2 species that can either diffuse to Mars' atmosphere or transported downward through Mars' subsurface. This cycle involves exothermic reactions that provide ample amounts of energy that can be then used for biosynthesis (e.g. nitrogen fixation) (Rubio & Ludden Reference Rubio and Ludden2005; Zehr et al. Reference Zehr, Jenkins, Short and Steward2005; Ducluzeau et al. Reference Ducluzeau, Lis, Duval, Schoepp-Cothenet, Russell and Nitschke2009). This allows for a quantitative comparison between using solely atmospheric nitrogen as a nutrient source and the maximum amount of subsurface nitrogen. Within this context, we also quantitatively assess the limiting factors for nitrogen fixation on Mars (e.g. available nitrogen, energy and water).

Results and discussion

Contradistinction to the Klingler hypothesis

Model estimates indicate that impacts, hydrodynamic escape and sputtering removed 99% of the Mars's initial atmosphere (McElroy et al. Reference McElroy, Kong and Yung1977; Bogard et al. Reference Bogard, Clayton, Marti, Owen and Turner2001). Earth scaling and models estimated an initial nitrogen atmospheric inventory ranging from 2 to 300 mbar (McKay & Stoker Reference McKay and Stoker1989), providing, at most, over three orders of magnitude more nitrogen for a primordial Martian biosphere. Taking into account Mars' current atmospheric nitrogen content (i.e. 0.2 mbar), this yields a maximum nitrogen missing sink of 2.8 mbar (9×1015 kg of N). A maximum nitrogen missing sink of 2.8 mbar is plausible as atmospheric nitrogen would accumulate over millennial time frames in the near-surface (i.e. top few metres) as water infiltration and leaching are minimal (Walvoord et al. Reference Walvoord2003; Graham et al. Reference Graham, Hirmas, Wood and Amrhein2008). Were the current entire atmospheric reservoir to be converted into biomass, Mars could at most support ∼1015 kg of biomass, or roughly 1 kg m−2 over the surface. By comparison, primary productivity (photosynthetic) on Earth is ∼1014 kg yr−1 (Field et al. Reference Field, Behrenfeld, Randerson and Falkowski1998). The terrestrial mass ratio of nitrogen in biomass to that in the atmosphere is ∼10−5 (Capone et al. Reference Capone, Popa, Flood and Nealson2006); if this ratio were to hold for Mars, we might expect ∼1010 kg of total biomass. Therefore, micro-organism growth could clearly not be supported if the only nitrogen source were from the atmosphere. However, if the 9×1015 kg of subsurface N were added to the mix, then biological processes could still be occurring (Klingler et al. Reference Klingler, Mancinelli and White1989). If just 10% of the atmospheric nitrogen lost between early Mars and contemporary Mars went into abiotic or biogenic sediments, from an initial nitrogen inventory of 18 mbar (Fox & Delgarno Reference Fox and Delgarno1983; Klingler et al. Reference Klingler, Mancinelli and White1989), then ∼1011 kg should be expected in the Martian rock record, comparable to the nitrogen sequestered in sediments on Earth, ∼1011 kg (Field et al. Reference Field, Behrenfeld, Randerson and Falkowski1998).

Activity on the Martian surface is limited to diffusion and heterogeneous photochemical processes. Lateral dissolution of atmospheric species would occur in thin liquid films with thicknesses ranging from ∼1 to 6 nm given atomic diameters are ∼1 to a few angstroms. Therefore, lateral dissolution of HNO3 would occur at Mars' surface. 1D gas-phase photochemical simulations, constrained by Viking NO measurements, made in the upper Martian atmosphere, show that HNO3 is one of the principal reservoir species for nitrogen in its lower atmosphere, representing ∼10−11 of Mars' current total nitrogen content. The Henry's law coefficients of other nitrogen oxides’ are orders of magnitude smaller than HNO3. For instance, the Henry's law coefficient for N2, NO, NO2, N2O, HNO2 and HNO3 are 6×10−4, 2×10−3, 1×10−2, 3×10−2, 49 and 2×105 M atm−1, respectively. Therefore, the dissolution of gas-phase HNO3 is extremely efficient. 1D photochemical simulations show that the current steady-state concentration of HNO3(g) just above the Martian surface is ∼5×105 molec. cm−3 (8×10−16 M of NO3), which represents ∼10−11 % of the total atmospheric nitrogen content, just above the Martian surface. Given that the total amount of thin liquid film ranges from 3.5×106 to 2×107 litres, 2×10−9 moles (1.5×10−10 kg) to 2×10−7 moles (1.5×10−8 kg) of NO3 is available at the Martian surface. Dissolution of N2 may also occur on slower timescales as its Henry's law constant is 6×10−4 M atm−1. The current steady-state concentration of N2(g) just above the Martian surface is ∼5×1015 molec. cm−3 (9.6×10−6 M N2), which equates to ∼10−6 % of the total atmospheric nitrogen content and ∼30 moles (1 kg) to ∼3000 moles (80 kg) of N2 (i.e. the amount of N2 that would be available at the Martian surface). Within these contexts, potential life would have to thrive under conditions of severe nitrogen deficiency as [HNO3] and [N2] in the Martian atmosphere is a small fraction of their respective amounts in Earth's atmosphere (e.g. ∼10−5 to 0.03, respectively, at the surface, Fig. 1).

Fig. 1. 1D model simulation of altitude (km) versus [HNO3(g)]Mars/[HNO3(g)]Earth and [N2(g)]Mars/[N2(g)]Earth.

Exothermic reactions as a source of energy for nitrogen fixation

Weiss et al. (Reference Weiss, Yung and Nealson2000) showed that subsurface Martian organisms could be supplied with a large energy flux from the oxidation of photochemically produced atmospheric H2 and CO diffusing into the regolith. Surface abundance measurements of these gases demonstrate that not more than a few percent of this available flux is actually consumed, suggesting that biological activity driven by H2 and CO is limited in the top few hundred metres of the subsurface, far exceeding the energy derivable from other atmospheric gases. They show for organisms at 30 m depth that the hydrothermal and chemical weathering energy is 2000 times more than previous estimates. This implies that the apparent scarcity of life on Mars is not attributable to lack of energy. The following thermodynamic calculations support this notion as exothermic reactions may occur on thin liquid films to release energy for nitrogen fixation.

Nitrogen is linked by a very strong triple bond and is energetically expensive (bond strength ∼945 kJ mole−1 – see reactions (1) and (2)). Fixed nitrogen (i.e. NH3, NH4+ and NOx or nitrogen that is chemically bound to either inorganic or organic molecules that can be released by hydrolysis to form NH3 and NH4+) is useful to living organisms (Ducluzeau et al. Reference Ducluzeau, Lis, Duval, Schoepp-Cothenet, Russell and Nitschke2009). Available quantities of free energy are crucial for supporting life and debated in terms of it being a requirement for nitrogen fixation:

(1)
$${\rm N}_{2({\rm g})} +{\rm 6H}^+ +{\rm 6e}^{\rm -} +{\rm energy} \to {\rm 2NH}_{\rm 3} {\rm.} $$
(2)
$${\rm N}_{2({\rm g})} +{\rm 8H}^+ +{\rm 8e}^{\rm -}+ {\rm energy} \to {\rm 2NH}_{\rm 3} +{\rm H}_{\rm 2} {\rm.} $$

For chemical reactions, we are often interested in the change in Gibbs free energy (ΔG),

(3)
$${\rm \Delta} G = {\rm \Delta} G^o + RT\ln \left( {\displaystyle{{(P_C )^m (P_D )^n} \over {(P_A )^x (P_B )^y}}}\right) = {\rm \Delta} G^o + RT\ln (Q),$$

where P is the pressure, T is the temperature, R is the universal gas constant, Q is the reaction quotient, A, B, C and D are reactants, respectively, m, n, x and y are coefficients of reactants, and products ΔGo is the standard free energy change (at 1 bar and 25°C). The standard free energies for nitrogen, oxygen, water and nitric acid in the following reaction

(4)
$$\displaystyle{1 \over 2}{\rm N}_{2({\rm g})} + \displaystyle{5 \over 4}{\rm O}_{2({\rm g})} + \displaystyle{1 \over 2}{\rm H}_2 {\rm O}_{{\rm (l)}} \Leftrightarrow {\rm HNO}_{3({\rm aq})} $$

are, respectively 0, 0, −236.65 and −110.88 kJ mole−1, yielding ΔGo=7.45 kJ mole−1. This reaction, requiring a positive change in Gibbs free energy, is not spontaneous and will proceed only with the addition of energy. We can now calculate conditions when the above reaction will proceed without the addition of free energy. Given Mars's mean atmospheric surface pressure of 5.53×10−3 atm (or 5.6 mbar), its N2(g) and O2(g) abundances are 2.7% and 0.13%, respectively, we approximate N2(g) and O2(g) pressures to be 1.49×10−4 and 7.19×10−6 atm, respectively. We assume here that water in the form of thin liquid films had unit activity. So, at equilibrium:

(5)
$${\rm \Delta} G = 0,$$
(6)
$${\rm \Delta} G^o = - RT\log (Q) = - 5.71\,\log (Q),$$
(7)
$$\log (Q) = \log \left( {\displaystyle{{a_{{\rm HNO}_3 _{({\rm aq})}}} \over {a_{{\rm N}_{\rm 2}} ^{1/2} a_{{\rm O}_{\rm 2}} ^{5/4}}}} \right) = - \displaystyle{{7.45} \over {5.71}} = - 1.304 \quad {\rm Or} \, Q = 0.0497,$$
(8)
$$a_{{\rm HNO}_{3({\rm aq})} } = Q \times (1.4 \times 10^{ - 4} )^{1/2} \times (7.19 \times 10^{ - 6} )^{5/4} = 2.00 \times 10^{ - 10}, $$

This value corresponds to a concentration of ∼3 nm (Lide Reference Lide2006). So, as long as [HNO3(aq.) or NO3] remains below ∼3 nm, reaction (2) will proceed with the liberation of free energy.

Given the fact that [HNO3(aq.) or NO3] is likely smaller than ∼3 nm (e.g. 8×10−16 M of steady-state HNO3 by 1D photochemical simulations) there is heat (not work) liberated for partial use as chemical energy. The more dilute the concentration at which nitrate is formed, the greater the amount of available work (not heat) will be available. If we assume that [HNO3] ranges from 10−16 to 10−9 M and the activity coefficient at this concentration is unity, the following calculations show that the free energy is

(9{\rm a})
$$\vskip-3pt Q = \displaystyle{{(10^{ - 9} )} \over {(1.4 \times 10^{ - 4} )^{1/2} \times (7.19 \times 10^{ - 6} )^{5/4} }} = 0.36,$$
(9{\rm b})
$$Q = \displaystyle{{(10^{ - 16} )} \over {(1.4 \times 10^{ - 4} )^{1/2} \times (7.19 \times 10^{ - 6} )^{5/4} }} = 3.61 \times 10^{ - 8}. $$

Then

(10)
$$\vskip3pt{\rm \Delta} G = {\rm \Delta} G^o + RT\,\log (Q) = 7.45 + 5.71( - 0.44) = 4.92\,{\rm kJ}\,{\rm mole}^{ - 1} \,{\rm and}\,{\rm \Delta} G = 7.45 + 5.71( - 7.44) = - 35.05\,{\rm kJ}\,{\rm mole}^{ - 1}. $$

These calculations show that the negative change in Gibbs free energy increases with decreasing HNO3. In other words, for the low concentration of HNO3 (in the form of NO3), we expect for pure water thin liquid films on the surface of Mars, reaction (2) is exergonic and may proceed spontaneously. Figure 2 exemplifies the relationship between the change in Gibbs free energy and HNO3 concentration. Temperatures shown correspond to the 186, 227 and 268 K. For the warmest temperature, HNO3 concentrations of <10−9 M yield a negative change in Gibbs free energy. At 186 K, concentrations of <3×10−11 M yield a negative change in Gibbs free energy. At the mean surface temperature of Mars (227 K), an HNO3 concentration of 8×10−11 M or less yields a negative change in Gibbs free energy.

Fig. 2. The change in Gibbs free energy (ΔG) versus HNO3 concentration at 186, 227 and 268 K.

The foregoing analyses fail to serve for anaerobic organisms, such as Clostridium, which can fix nitrogen in the absence of oxygen. The production of aqueous ammonia from nitrogen and hydrogen is attended by ΔGo=−26 and ΔHo=−84.84 kJ mole−1 (reaction 4).

(11)
$$\vskip3pt\displaystyle{1 \over 2}{\rm N}_{2({\rm g})} + \displaystyle{3 \over 2}{\rm H}_{2({\rm g})} + {\rm H}_{\rm 2} {\rm O}_{({\rm l})} \Leftrightarrow {\rm NH}_{\rm 4} {\rm OH}_{4({\rm aq)}}. $$

In comparison, if oxygen were present (reaction 11), replacing water (reaction 11), ΔGo=−263.01 kJ mole−1.

(12)
$$\vskip3pt\displaystyle{1 \over 2}{\rm N}_{2({\rm g})} + \displaystyle{5 \over 2}{\rm H}_{2({\rm g})} + \displaystyle{1 \over 2}{\rm O}_{2({\rm g})} \Leftrightarrow {\rm NH}_{\rm 4} {\rm OH}_{4({\rm aq)}}. $$

An analogous assessment can be made between reactions (2) and (12), where ΔGo=−110.88 and ΔHo=−205.44 kJ mole−1 in reaction (12) due to the presence of hydrogen

(13)
$$\vskip3pt\displaystyle{1 \over 2}{\rm N}_{2({\rm g})} + \displaystyle{1 \over 2}{\rm H}_{2({\rm g})} + \displaystyle{3 \over 2}{\rm O}_{2({\rm g})} \Leftrightarrow {\rm HNO}_{{\rm 3(aq)}}. $$

Photochemical reactions as a source of energy for nitrogen fixation

Other forms of energy, such as UV radiation at the Martian surface can also produce fixed nitrogen species, such as NO(g) and NO2(g) via photolysis of HNO3(g) and NO3 (aq.), which has been investigated extensively in the lab (Dubowski et al. Reference Dubowski, Colussi, Boxe and Hoffmann2002; Boxe et al. Reference Boxe2003, Reference Boxe2005,  Reference Boxe2006) and to a limited extent via multiphase modelling (Boxe & Saiz-Lopez Reference Boxe and Saiz-Lopez2008). The gas phase concentration of HNO3 at the surface of Mars is ∼5×105 molec cm−3, which is equivalent to 8×10−16 M of NO3. Since Boxe & Saiz-Lopez (Reference Boxe and Saiz-Lopez2008) calculated a 3×104 concentration enhancement from 9 μM NO3 in a 300-nm-thick liquid layer, we calculate a 4.5×105 concentration effect for 8×10−16 M of NO3 absorbed in a 6-nm-thick interfacial film. We then initialize our model with [NO3]=4×10−10 M. The actinic flux spectrum at the surface of Mars is from 200⩽λ/nm⩽400 (Yung & Demore Reference Yung and Demore1999) so the condensed phase model incorporates the predominant reactions representative of nitrate photochemistry in this wavelength region. The actinic flux spectrum for Mars' was also weighted to take into account NO3 absorption bands centred at 201 nm (εmax=9500 M−1 cm−1) and 302 nm (εmax=7.14 M−1 cm−1) and NO2 absorption bands centred at 220 nm (εmax=5800 M−1 cm−1), 318 nm (εmax=10.90 M−1 cm−1) and 354 nm (εmax=22.90 M−1 cm−1). A volumetric factor was quantified by taking the average of the upper and lower limit reaction rate enhancement factors (unit-less numbers), obtained in the laboratory by Grannas et al. (Reference Grannas, Shepson and Filley2004) and Takenaka et al. (Reference Takenaka1996), 40 and 2.4×103, respectively, yielding 1.22×103. Therefore, the reaction rates are quantified by incorporating a volumetric factor, volumetric. Table 1 lists the major reactions pertaining to nitrate photochemistry, their condensed phase reaction rates and their interfacial reaction rates. Although we scale the secondary and tertiary non-photolytic reaction rates to give a thorough representation of nitrate photochemistry on the Martian surface, as shown by Boxe & Saiz-Lopez (Reference Boxe and Saiz-Lopez2008), the photolytic reactions dominate the evolution of NO, NO2 and NO2.

Table 1. Interfacial film reactions and rate constants

a Aqueous phase reaction rate constants were obtained from Mack & Bolton (Reference Mack and Bolton1999).

b QLL rate reaction rate constants were quantified by including the ‘volumetric’ factor (Grannas et al. (Reference Grannas, Bausch and Mahanna2007), Takenaka et al. (1996)).

d $J_{{\rm NO}_{\rm 2} ^{\,\rm -}} $ was extrapolated from Zuo & Deng (Reference Zuo and Deng1998).

e Volumetric ∼8.20×10−4 (Grannas et al. (2007); Takenaka et al. (1996)).

Figure 3 shows a diurnal profile of NO, NO2 and NO2 evolution in a 20-nm-thick interfacial film due to heterogeneous photochemical processing of NO3. NO, NO2 and NO2 exhibit maximum concentrations of 1.3×102, 2×103 and 9.4×103 molec cm3. Although the total integrated flux over 200–400 nm at the surface of Mars is comparable to the Earth's, the shorter wavelengths contribute a much greater proportion of this UV flux (e.g. 200⩽λ/nm⩽315). This is, the fact that NO2 is a primary photolytic product, due to its greater branching ratio (compared to NO2) for photolytic production combined with the fact that NO3 centred at 201 nm has a greater absorption band, compared to the NO2 absorption band centred at 220 nm explains the greater production of NO2, compared to NO. In addition, the decrease in contribution of the actinic flux spectrum at λ>315 nm also helps to explain the smaller amount of NO produced in the thin liquid film since nitrite's absorption bands are centred at 318 and 354 nm. The concentration of NO3 in the interfacial film is constant due to the small quantum yield for NO2 and nitrite production and the fact that NO3 is replenished by atmospheric deposition of HNO3. This phenomenon implies that Martian surface is not a permanent sink for atmospheric nitrogen (Fig. 3).

Fig. 3. Simulated evolution as a function of time (hours) and concentration molec cm3 of NO, NO2 and NO2 during the heterogeneous photochemical processing of NO3 in thin liquid films on the Martian surface.

Downward transport of nitrogen (HNO3(g) and N2(g))

Figure 4 also shows that there will be downward transport of nitrogen species to the Martian subsurface. In desert environments on Earth, water infiltration and leaching are minimal (Walvoord et al. Reference Walvoord2003; Graham et al. Reference Graham, Hirmas, Wood and Amrhein2008) so the downward transport of atmospheric nitrogen will be dominated by gas-phase diffusion. Because the mean-free paths of HNO3 and N2 through the background atmosphere are greater than the typical pore size (∼1 μm), molecular collisions with the walls of the pores dominate the transport of these molecules. This is known as Knudsen diffusion, for which we can take D∼(2εro/3τ)(2kT/πm)1/2, where ro is the pore size, ε is the porosity, τ is the tortuosity and m is the molecular mass of the diffusing component (Bullock et al. Reference Bullock, Stoker, McKay and Zent1994; Harman & McKay Reference Harman and McKay1995). Consistent with diffusion models of Mars, we use ro=6×10−4 cm, ε=0.5 and τ∼5 for depths less than ∼1 km (Squyres et al. Reference Squyres, Clifford, Kuzmin, Zimbelman, Costard, Kieffer, Jakosky, Snyder and Matthews1992; Mellon & Jakosky Reference Mellon and Jakosky1993; Bullock et al. Reference Bullock, Stoker, McKay and Zent1994; Harman & McKay Reference Harman and McKay1995). At a mean surface temperature of 220 K, we find that, above 1 km, $D_{{\rm HNO}_3} $ ∼0.54 cm2 s−1 and $D_{{\rm N}_{\rm 2}} $∼0.81 cm2 s−1. The maximum sink that could result from the Knudsen diffusion can be estimated by the rate of downward diffusion, F (molecules cm−2 s−1), specified by F=−D(dn/dz), where n is the number density of the diffusing component (molecules cm−3), z is the depth (cm) with the surface set to z=0, D is the diffusion coefficient (cm−2 s−1) and dn/dz is the density gradient (molecules cm−4). 1D photochemical simulations show that the number density of HNO3 and N2 at Mars' surface is 1.0×105 and 5.81×1015 molecules cm−3, respectively. Thus a putative biotic layer at 10 m depth would produce HNO3 and N2 sinks of −54 and −5×1012 molecules cm−2 s−1, respectively, which is an ample supply of available nitrogen that can be efficiently transported to the subsurface to support a putative subsurface biosphere.

Fig. 4. Simplified schematic illustrating HNO3(g) deposition and absorption in thin liquid films on the Martian surface (i.e. regolith (a) and ice (b)), followed by heterogeneous photochemical processing and transport of NO3, NO2, NO2 and NO to lower depths via mobility of thin liquid films to the Martian subsurface.

Conclusions and implications

Mars' total nitrogen atmospheric nitrogen content is 0.2 mbar (6.5×1014 kg). The lower-limit nitrogen content in Martian biomass is 1010 kg. These constraints are indicative of conditions of severe nitrogen deficiency. Over time, Mars' subsurface might have accumulated 2.8 mbar (9×1015 kg of N), which indicates that biological metabolism would not be inhibited at Mars' subsurface as suggested by the Klingler hypothesis (Klingler et al. Reference Klingler, Mancinelli and White1989).

We show for the first time that the negative change in Gibbs free energy increases with decreasing HNO3(g) (NO3(aq)) in metastable thin liquid films that may exist at the surface of Mars, where additional reaction pathways are exergonic and may proceed spontaneously, thus providing an ample source of energy for nitrogen fixation on Mars.

Condensed phase simulations show that photochemically produced fixed nitrogen species are not only released from the Martian surface to the gas-phase, but more importantly, transported to lower depths from the Martian surface in transient thin liquid films. For example, a putative biotic layer at 10 m depth would produce HNO3 and N2 sinks of −54 and −5×1012 molecules cm−2 s−1, respectively. This is an ample supply of available nitrogen that can be efficiently transported to the subsurface. The downward transport and release to the atmosphere of photochemically produced fixed nitrogen species suggests the existence of a transient nitrogen cycle on Mars.

References

Bogard, D.D., Clayton, R.N., Marti, K., Owen, T. & Turner, G. (2001). Space. Sci. Rev. 96, 425458.CrossRefGoogle Scholar
Boxe, C.S. & Saiz-Lopez, A. (2008). Atmos. Chem. Phys. 8, 48554864.CrossRefGoogle Scholar
Boxe, C.S. et al. (2003). J. Phys. Chem. A 107, 1140911413.CrossRefGoogle Scholar
Boxe, C.S. et al. (2005). J. Phys. Chem. A 109, 85208525.CrossRefGoogle Scholar
Boxe, C.S. et al. (2006). J. Phys. Chem. A 110, 76137616.CrossRefGoogle Scholar
Boxe, C.S. et al. (2011). Int. J. Astrobiol. submitted.Google Scholar
Bullock, M., Stoker, C., McKay, C. & Zent, A. (1994). Icarus 107, 142154.CrossRefGoogle Scholar
Capone, D.G., Popa, R., Flood, B. & Nealson, K. (2006). Science 312, 708709.CrossRefGoogle Scholar
Dubowski, Y., Colussi, A.J., Boxe, C.S. & Hoffmann, M.R. (2002). J. Phys. Chem. A 106, 69676971.CrossRefGoogle Scholar
Ducluzeau, A.-L., Lis, R.V., Duval, S., Schoepp-Cothenet, B., Russell, M. & Nitschke, W. (2009). Trends Biochem. Sci. 34, 915.CrossRefGoogle Scholar
Field, C.B., Behrenfeld, M.J., Randerson, J.T. & Falkowski, P. (1998). Science 281, 237240.CrossRefGoogle Scholar
Fox, J.L. & Delgarno, A. (1983). J. Geophys. Res. 88(A11), 90279032.CrossRefGoogle Scholar
Graham, R.C., Hirmas, D.R., Wood, Y.A. & Amrhein, C. (2008). Geology 36, 259262.CrossRefGoogle Scholar
Grannas, A.M., Bausch, A.R. & Mahanna, K.M. (2007). J. Phys. Chem. A 111, 1104311049.CrossRefGoogle Scholar
Grannas, A.M., Shepson, P.B. & Filley, T.R. (2004). Global Biogeochem. Cycles 18, doi: 10.1029/2003GB002133, GB1006(1-10).CrossRefGoogle Scholar
Harman, H. & McKay, C. (1995). Planet. Space Sci. 43, 123128.CrossRefGoogle Scholar
Huber, H. et al. (2002). Nature 417, 6367.CrossRefGoogle Scholar
Jakosky, B.M. & Phillips, R.L. (2001). Nature 412, 237244.CrossRefGoogle Scholar
King, M.D., France, J.L., Fisher, F.N. & Beine, H.J. (2005). J. Photochem. Photobiol. A 176, 3949.CrossRefGoogle Scholar
Klingler, J.M., Mancinelli, R.L. & White, M.R. (1989). Adv. Space Res. 9, 173176.CrossRefGoogle Scholar
Lide, David R. (eds) (2006). CRC Handbook of Chemistry and Physics, 89th edn (Internet Version 2009). CRC Press/Taylor and Francis, Boca Raton, FL.Google Scholar
Mack, J. & Bolton, J.R. (1999). J. Photochem. and Photobiol. A 128, 113.CrossRefGoogle Scholar
Mancinelli, R.L. & Banin, A. (2003). Int. J. Astrobiol. 2, 217225.CrossRefGoogle Scholar
McElroy, M.B., Kong, T.Y. & Yung, Y.L. (1977). J. Geophys. Res. 82, 43794388.CrossRefGoogle Scholar
McKay, C.P. & Stoker, C.R. (1989). Rev. Geophys. 27, 189214.CrossRefGoogle Scholar
Mellon, M. & Jakosky, B. (1993). J. Geophys. Res. 98, 33433364.Google Scholar
Price, P.B. (2007). Microbiol. Ecol. 59, 217231.CrossRefGoogle Scholar
Qiu, R., Green, S.A., Honrath, R.E., Peterson, M.C., Lu, Y. & Dziobak, M. (2002). Atmos. Environ. 36, 25632571.CrossRefGoogle Scholar
Rubio, L.M. & Ludden, P.W. (2005). J. Bacteriol. 187, 405414.CrossRefGoogle Scholar
Squyres, S.W., Clifford, S.M., Kuzmin, R.O., Zimbelman, J.R. & Costard, F.M. (1992). In Mars, ed. Kieffer, H., Jakosky, B., Snyder, C. & Matthews, M., pp. 523554. University of Arizona Press, Tuscon, AZ.Google Scholar
Summers, D.P. & Khare, B. (2007). Astrobiology 7, 333, doi: 10.1089/ast.2006.0032.CrossRefGoogle Scholar
Takenaka, N. et al. (1996). J. Phys. Chem. 100, 1387413884.CrossRefGoogle Scholar
Walvoord, M.A. et al. (2003). Science 302, 10211024.CrossRefGoogle Scholar
Weiss, B.P., Yung, Y.L. & Nealson, K.H. (2000). Proc. Natl Acad. Sci. USA 97, 13951399.CrossRefGoogle Scholar
Yung, Y.L. & Demore, W.B. (1999). Photochemistry of Planetary Atmospheres, Oxford University Press, Oxford.CrossRefGoogle Scholar
Zehr, J.P., Jenkins, B.D., Short, M.D. & Steward, G.F. (2005). Environ. Microb. 5, 539554.CrossRefGoogle Scholar
Zuo, Y. & Deng, Y. (1998). Chemosphere 36, 181188.CrossRefGoogle Scholar
Figure 0

Fig. 1. 1D model simulation of altitude (km) versus [HNO3(g)]Mars/[HNO3(g)]Earth and [N2(g)]Mars/[N2(g)]Earth.

Figure 1

Fig. 2. The change in Gibbs free energy (ΔG) versus HNO3 concentration at 186, 227 and 268 K.

Figure 2

Table 1. Interfacial film reactions and rate constants

Figure 3

Fig. 3. Simulated evolution as a function of time (hours) and concentration molec cm3 of NO, NO2 and NO2 during the heterogeneous photochemical processing of NO3 in thin liquid films on the Martian surface.

Figure 4

Fig. 4. Simplified schematic illustrating HNO3(g) deposition and absorption in thin liquid films on the Martian surface (i.e. regolith (a) and ice (b)), followed by heterogeneous photochemical processing and transport of NO3, NO2, NO2 and NO to lower depths via mobility of thin liquid films to the Martian subsurface.