You are on page 1of 13

Applied Geochemistry 143 (2022) 105373

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Microbial vs abiotic origin of methane in continental serpentinized


ultramafic rocks: A critical review and the need of a holistic approach
Giuseppe Etiope a, b, *, Christopher Oze c
a
Istituto Nazionale di Geofisica e Vulcanologia, Sezione Roma 2, Rome, Italy
b
Faculty of Environmental Science and Engineering, Babes-Bolyai University, Cluj-Napoca, Romania
c
Occidental College, Los Angeles, United States

A R T I C L E I N F O A B S T R A C T

Editorial handling by Dr. Xianjun Xie Continental ultramafic rock systems, through the process of serpentinization, provide chemical and biochemical
pathways that lead to the production of methane. The extent to which rock-water-gas reactions and organisms
Keywords: supply methane in these systems is a matter of considerable discussion and debate. Deciphering the interplay of
Abiotic methane abiotic and microbial methane observed at the surface requires several lines of reasoning as well as a variety of
CO2 hydrogenation
analyses. Despite using multiple models and interpretative tools, conclusions for the origin of methane at a
Fluid inclusions
particular site may vary or diverge from regional or global observations. Here, we critically address how possible
Methanogens
Serpentinization conclusions of microbial versus abiotic methane in continental serpentinization systems may be interpreted and
Peridotite reinterpreted. We review fundamental concepts, advantages and limits, for three major methane origin models:
Hyperalkaline water (a) abiotic CO2 hydrogenation supplying gas reservoirs, (b) derivation from fluid inclusions in olivine-rich rocks,
Ophiolite and (c) microbialgenesis in aquifers. We use the case of methane in the Samail ophiolite of Oman as an
emblematic example of multiple interpretations; we identify ambiguous information offered by methane clum­
ped isotopes and molecular gas compositions (e.g., the meaning of gaseous hydrocarbons heavier than methane),
and suggest key tools, such as radiocarbon (14C) in methane, which may solve interpretative issues. The major
constraint in any model of methane origin is the capability to sustain continuous gas flows, in terms of methane
emission intensity, longevity and spatial extension, such as in natural gas sedimentary systems. Overall, this
review suggests that any site interpretation can benefit from a holistic approach, integrating geochemical,
geological and biological data with gas flow dynamics, as well as including regional and global contextualization.

1. Introduction methane-bearing peridotite systems are the subject of a wide variety of


studies in planetary geology, subsurface microbiology and astrobiology
Methane (CH4), often associated with heavier hydrocarbons (ethane, (e.g., Oze and Sharma, 2005; Schrenk et al., 2013; Etiope and Sherwood
C2H6; propane, C3H8), has been reported in aquifers, springs and gas Lollar, 2013 and references therein). The exact origin of such methane
vents in continental serpentinized ultramafic rock systems (ophiolites, is, however, the subject of current discussion and debate, and three main
peridotite massifs and intrusions) worldwide (an updated list of inves­ alternative or complementary theories have been proposed by different
tigated sites in 20 countries is reported in Table S1). Ultramafic rocks are research groups (Fig. 1): (a) Abiotic methane from low-medium tem­
devoid of ancient organic matter, thus, methane is generated by pro­ perature (<150 ◦ C), Fischer-Tropsch Type (FTT) reactions, specifically
cesses that are different than those producing traditional natural gas in CO2 hydrogenation (or Sabatier reaction) in fractures and subsequent
sedimentary and petroliferous rocks. In sedimentary rocks methane is gas migration and accumulation in reservoirs. This model includes the
derived from ancient life forms, i.e., microbial and thermal decompo­ hypothesis of chromitites as source rocks (Etiope et al., 2018; Etiope and
sition of organic matter. In ultramafic rock systems methane can be a Whiticar, 2019) and the importance of metals for catalysis. Hereafter,
product of chemical reactions and an energy source for life. In active the term “CO2 hydrogenation-reservoir model” is used.
serpentinization systems, methane can also be derived from living (b) Abiotic methane from fluid inclusions in olivine-rich rocks,
chemosynthetic microbial communities (methanogens). Hence, originated from higher temperatures FTT synthesis after magmatic fluid

* Corresponding author. Istituto Nazionale di Geofisica e Vulcanologia, Sezione Roma 2, Rome, Italy.
E-mail address: giuseppe.etiope@ingv.it (G. Etiope).

https://doi.org/10.1016/j.apgeochem.2022.105373
Received 9 February 2022; Received in revised form 31 May 2022; Accepted 10 June 2022
Available online 15 June 2022
0883-2927/© 2022 Elsevier Ltd. All rights reserved.
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Fig. 1. A scheme summarizing the three main models


of methane origin in continental serpentinized ultra­
mafic rock systems. Each model must be capable of
generating enough gas amounts and pressures to
sustain the continuous and long-lasting emissions
observed at the surface, including bubble plumes in
springs or surface pools (a), methane in solution
transported by groundwater (b), and dry gas seeps
(c). Photos by G. Etiope: (a) bubbling in one of the
several hyperalkaline springs in the Ronda ophiolite;
(b) CH4-enriched water pumped in the Terme di
Genova spa; (c) flames from the large Chimaera
burning seep in Turkey. Additional photos are re­
ported in Fig. S4.

cooling (Klein et al., 2019; Grozeva et al., 2020). radiocarbon (14C) CH4 analyses, only occasionally performed in ser­
(c) Microbial origin in aquifers, through the action of hydro­ pentinized ultramafic rock environments; (b) the attribution of a mi­
genotrophic (using CO2), formatotrophic (using formate), and/or ace­ crobial origin by CH4 clumped isotopes; (c) the interpretation of ethane
totrophic (using acetate) methanogens (e.g., Miller et al., 2018; Nothaft and propane data; (d) gas flow dynamics constraints, whereby, any
et al., 2021). methane origin, to be acceptable, shall be capable of generating,
Ultramafic rock systems can also receive variable inputs of thermo­ continuously and over long-time scales, enough gas to produce those
genic gas from underlying or adjacent sedimentary rocks. This compo­ pressure gradients and gas flows compatible with the relevant occur­
nent, however, has generally been considered accessory (e.g., Etiope rences and emissions observed at the surface in many places. We show
et al., 2011a; Szponar et al., 2013). Of course, abiotic methane can also that a given combination of geochemical data can be explained by
be produced by other mechanisms, as reviewed in Etiope and Sherwood multiple hypotheses, but critical and dynamic relationships that may
Lollar (2013), which refer to temperature and pressure conditions that solve interpretative problems are identified. Points (a) to (d) reveal the
are much higher than those of continental ultramafic rock systems. need of a holistic interpretative approach, capable of incorporating
These include the interaction between olivine hydration and dissolved C multiple factors and processes connected to what is specifically observed
in hydrothermal systems (e.g., see a review of the mechanisms in in a given site, but that operate on a wider spatial scale, both regional
McCollom, 2016, and a review of the hydrothermal FTT laboratory ex­ and global.
periments in Huang et al., 2021), or the reaction beween H2 and
elemental C (graphite; e.g., Deville and Prinzhofer, 2016), which is 2. Three models of methane origin in continental serpentinized
theoretically applicable to deep, high T and P systems (French, 1966), ultramafic rocks
such as subducted slabs (Brovarone et al., 2017). Continental ultramafic
rocks, especially those in ophiolites (with a few exceptions of ophiolites 2.1. Abiotic origin related to moderate temperature gas-phase CO2
in high heat flow regions), do not refer to present-day high temperature hydrogenation and accumulation in reservoirs
hydrothermal systems, as indicated by the generally low geothermal
gradients, temperatures in deep boreholes crossing the whole ophiolite The pioneering studies of Abrajano et al. (1990), Lyon et al. (1990)
sequence (e.g., Demirel and Günay, 2000) and modern serpentinization and Fritz et al. (1992) on gas seeps in the Philippines, New Zealand and
temperatures (e.g., Ellison et al., 2021). High temperature hydrothermal Oman ophiolites, and more recent research in Europe, North America,
processes can be invoked for methane formed within the Japan, New Caledonia and again Oman (e.g., Etiope et al. 2011a; 2013a,
mafic-ultramafic rocks before their emplacement on continents, and 2013b, 2016, 2017; Szponar et al., 2013; Suda et al., 2017; D’Alessandro
preserved in inclusions, as considered in the fluid inclusion model et al., 2018; Vacquand et al., 2018), suggest a dominant abiotic origin of
mentioned above. Methane of pure magmatic origin is also excluded by methane (i.e., the gas is not the direct result of microbes or remineral­
multiple lines of evidence, including the relatively low temperatures of ization of organic matter). Variable, but minor biotic (thermogenic or
methane formation revealed by methane clumped analyses and other microbial) components can be present. The abiotic origin refers mainly
geo-thermometers (e.g., Etiope and Sherwood Lollar, 2013; Wang et al., to FTT synthesis, mostly low temperature CO2 hydrogenation or Sabatier
2015; Blank et al., 2017; Young et al., 2017; Suda et al., 2022). reaction (CO2 + 4H2 = CH4 + 2H2O) occurring in fractured rocks.
Here, advantages and limits of the several models and interpretative Although some works consider FTT synthesis inefficient (e.g., McCol­
tools are examined. A specific discussion develops around the lom, 2013; Barbier et al., 2020), this conclusion should be restricted to
emblematic case of the gas observed in the Samail ophiolite (Oman) by aqueous CO2 hydrogenation. It is well known that the Sabatier reaction
several authors who suggested different interpretations (e.g., Neal and involves heterogeneous catalysis and operates only in dry, gas-phase
Stanger, 1983; Fritz et al., 1992; Vacquand et al., 2018; Nothaft et al., conditions, while aqueous FTT synthesis can proceed only through ho­
2021). Reflections presented here focus on: (a) the importance of mogeneous catalysis, with dissolved catalysts and synthetic ligands

2
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Fig. 2. Stable carbon and hydrogen isotope ratios of


methane in continental ultramafic rock systems
(ophiolites and peridotite massifs). Redrawn from
Etiope and Whiticar (2019; and references therein;
see also references in Table S1), with the addition of a
burning CH4–H2 rich gas seep in Indonesia (Tanjung
Api; East Sulawesi Ophiolite; Subroto et al., 2004),
the Oman data from Nothaft et al. (2021; only sam­
ples analysed for paired clumped-isotopes are added)
and Precambrian shield data considered dominantly
abiotic (grey crosses; Sherwood Lollar et al., 2006).
Small grey dots refer to biotic (thermogenic and mi­
crobial) gas from sedimentary and petroleum systems
(dataset from Milkov and Etiope, 2018). A similar
diagram including geothermal-volcanic gas and gas in
igneous inclusions is reported in Etiope and Sherwood
Lollar (2013).

(Etiope and Whiticar, 2019 and references therein). Laboratory experi­ proportion of methane was proposed to originate in chromitites as they
ments using or promoting gas-phase CO2 hydrogenation (strictly showed, among the whole ophiolite sequences in Greece and Brazil, to
following Sabatier’s rules) obtained substantial and unambiguous be the only rocks with relevant concentrations of CH4, H2 and potential
amounts of methane over a wide range of temperatures (e.g., Etiope and metal Sabatier catalysts (Etiope et al., 2018; De Melo Portella et al.,
Ionescu, 2015; Barbier et al., 2020 and references therein). A more 2019). Methane observed in the chromitite microfractures (not in in­
detailed discussion on FTT reactions (with a distinction of CO-based, clusions) has the same isotopic composition of surface seeps (Etiope
Fischer-Tropsch reaction in sensu strictu) and their occurrence in natu­ et al., 2018). It is then hypothesized that the gas progressively migrates
ral geological systems (gas phase and aqueous solutions) is in Etiope and and accumulates in permeable reservoir rocks, in a
Whiticar (2019). The interpretation of this origin was typically based on source-accumulation-seepage system conceptually similar to natural gas
multiple lines of evidence, primarily focusing on stable C and H isotopes systems in sedimentary basins (Etiope and Whiticar, 2019). This model
composition of CH4 (δ13C and δ2H; Fig. 2), combined with the molecular is mainly limited by the poorly known effectiveness and abundance of
gas composition, including H2 and its isotopic composition, preliminary minerals that can act as catalysts (e.g., chromium, ruthenium, iron
clumped-isotope analyses (either from surface dry gas seeps or gas dis­ minerals) that can support FTT reactions over long-time scales at mod­
solved in groundwater), the geological-mineralogical context and other erate temperature. Although Sabatier reaction based either on artifi­
geochemical proxies when possible (Etiope and Sherwood Lollar, 2013; cially prepared Ru-catalysts or natural chromitite samples has been
Etiope and Whiticar, 2019 and references therein). In particular, a demonstrated (Etiope and Ionescu, 2015; Ueda et al., 2021), studies
addressing how metals in natural rocks (generally as oxides or sulfides)
can work as Sabatier catalysts are lacking: can ruthenium minerals (e.g.,
laurite), dispersed at low concentrations (ppb to ppm levels) in the
chromitite, be sufficient to promote methane formation, and what would
be their support (typical of catalysts in Sabatier reactions)? Or is chro­
mium, which is more abundant in ultramafic rocks, a better candidate as
catalyst? If methane is produced at temperatures below 150 ◦ C, as
clumped-isotopes (and geothermal gradients) suggest in several cases,
then the catalyst must be operative at such low temperatures. We know
that ruthenium is effective below 150–200 ◦ C, while other catalysts,
based on Ni and Fe, require temperatures above 200 ◦ C (Etiope and
Whiticar, 2019 and references therein). Chromium is a less studied
catalyst, and we cannot exclude that it can work at low temperatures
over long time scales. It is interesting to note, however, that chromitites
have mineralogical signatures indicative of (a) H2 flow, such as laurite
alteration by desulfurization along microfractures (Garuti and Zaccarini,
1997), and (b) potential CO2-hydrogenation, such as the presence of
amorphous carbonaceous material, typically formed during the reaction
on the catalyst (e.g., Baraj et al., 2016), as observed in some Ru-rich
Fig. 3. Δ13CH3D vs Δ12CH2D2 plot. Continental ultramafic rock systems and chromitites in Greece (Economou-Eliopoulos et al., 2019). H2 flow is
Precambrian shield data are from Young et al. (2017), Blank et al. (2017) and also necessary to remove reaction residues and regenerate the catalyst,
Nothaft et al. (2021). Vertical grey dashed lines are continental ultramafic thereby, ensuring its activity for prolonged times (e.g., Forzatti and
systems where only Δ13CH3D was measured (data from Wang et al., 2015; Lietti, 1999). The available data are, however, scarce; this is a research
Woycheese et al., 2017; Cumming et al., 2019; Suda et al., 2022). Submarine
area that requires more exploration. CO2 hydrogenation is widely
hydrothermal fields data are from Labidi et al. (2020). Laboratory abiotic gas
studied in chemical catalysis, where several metals are tested (e.g., see
data are from Young et al. (2017) and Zhang et al. (2021). Blue dashed line
encompasses microbial data (microbialgenesis zone) from Young (2020). (For
review in Wang et al. (2011)), but for improving our understanding of
interpretation of the references to colour in this figure legend, the reader is the Sabatier synthesis in natural rocks, further studies are required.
referred to the Web version of this article.)

3
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Fig. 4. A scheme summarizing pros, cons and unresolved questions for the three main models of methane origin in continental serpentinized ultramafic rock systems.

2.2. Abiotic origin related to fluid inclusions in olivine-rich rocks producing CH4 (methanogens), suggesting a frequent, dominant mi­
crobial origin for the gas when it is associated to hyperalkaline (pH > 9)
Another proposed abiotic origin of methane observed in surface waters (e.g., Brazelton et al., 2017; Miller et al., 2018; Nothaft et al.,
seeps and springs refers to fluid inclusions in olivine-rich rocks, with 2021). Methanogens have been detected in some cases (The Cedars in
methane originated at higher temperature (>200–300 ◦ C), via FTT California, Voltri in Italy, Samail in Oman and Chimaera seep in Turkey;
synthesis after magmatic fluids cooling (Klein et al., 2019; Grozeva Kohl et al., 2016; Brazelton et al., 2017; Miller et al., 2018; Zwicker
et al., 2020). Hydrocarbons leaching from the fluid inclusions would et al., 2018; Nothaft et al., 2021), but not in others (Cabeco de Vide in
contribute significantly to surface gas emissions. This model is suggested Portugal; Tiago and Veríssimo, 2013; Tablelands in Canada; Morrill
by the discovery of CH4-rich fluid inclusions in peridotite and dunite et al., 2014; CROMO in California; Twing et al., 2017; Hakuba-Happo in
from ophiolites in Italy, Western United States and the Philippines (Klein Japan; Suda et al., 2022). A dominant microbial component is suggested
et al., 2019), and by the similitude in gas isotopic composition between by 13C-depleted CH4 values (δ13C < − 50‰) in some hyperalkaline
inclusions and surface seeps in the Philippines (Grozeva et al., 2020). springs (Bosnia, California, Spain; Fig. 2). New microbiological studies
This hypothesis is analogous to the one of McDermott et al. (2015) and are proposing that methane can be microbial also when it is
13
Wang et al. (2018) for methane in submarine hot springs. Its limits are C-enriched (even with positive δ13C values), as in Oman, due to
related to gas flow dynamics constraints, mainly the lack of a model that C-limited conditions in peridotites (Nothaft et al., 2021). A more general
can explain how large volumes of inclusion-rich rocks can be involved in microbiological hypothesis, dismissing abiotic theories in general, is
the gas leaching process to assure long-lasting and continuous gas flow proposed by Xia and Gao (2021), for which the interpretation of an
to the surface. Although Klein et al. (2019) hypothesized a process of abiotic origin is acceptable only for methane in fluid inclusions in some
fluid inclusion cooling and progressive formation of H2, it is not clear mantle minerals. Besides the observation that methanogens have not
how fluid inclusions in a given volume of rock, may have, individually always been detected in ultramafic rock aquifers or springs, a major
and simultaneously, all the ingredients in the right proportions neces­ limit of the microbial model refers to the fossil, radiocarbon (14C) free
sary for CH4 production (H2 ≫ CO2 and catalyst). Available nature of methane analysed so far, while C in the hyperalkaline water,
clumped-isotope analyses of gas discharged at the surface (Fig. 3) also where methanogens are detected, has measurable 14C quantities; thus,
suggest that methane is formed at temperatures below those typically methane cannot be originated from C in that water. A model explaining
considered for gas production in fluid inclusions. It is important to how microbes may use a 14C-free substrate is missing. The study of
outline, however, that many CH4-bearing inclusions observed in ultra­ endolithic microbial activity within deep mafic-ultramafic rocks may be
mafic rocks (including chromitites) are secondary and originated by a key approach (e.g., Fones et al., 2022), but it should be associated with
healing of fractures (Melcher et al., 1997; Van den Kerkhof and Hein, the analysis of methane, and its 14C content, occluded in the rocks. In
2001) implying that the gas was initially free along fractures, originated addition, CH4 production in water appears incompatible with CH4–H2O
outside the inclusion, and then trapped within the inclusion. The trap­ disequilibrium generally observed in the hyperalkaline waters (Etiope
ping could take place after the FTT reaction occurred (as in Salvi and and Whiticar, 2019 and references therein). As for the fluid inclusion
Williams-Jones, 1997), so there would not be a conflict between the CO2 hypothesis, a model is necessary to describe how microbes in aquifers
hydrogenation-reservoir model and the inclusions’ model. Klein et al. can sustain the pervasive, long-lasting and intense gas flows observed at
(2019) conceived fluid trapping at high temperatures (>400 ◦ C), before the surface in many places (Fig. 1 and Fig. S4). The frequent presence of
H2 formation. heavier hydrocarbons (ethane to butane) indicates that microbialgenesis
must be systematically complemented by other hydrocarbon producing
processes.
2.3. The microbial origin
Pros, cons and unresolved questions for the three main models of
methane origin are summarised in Fig. 4.
Recent microbiological studies have emphasized the role of microbes

4
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

3. Unravelling biotic vs abiotic CH4 origin by isotopes Beyond Oman, paired (Δ12CH2D2 and Δ13CH3D) clumped isotope
analyses are today available from four sites of serpentinized ultramafic
13
3.1. The stable “bulk” isotope ratios C/12C and 2H/H of CH4 rocks, in Italy, Portugal, Turkey (Young et al., 2017) and California
(Blank et al., 2017); from five other sites (in California, Canada,
The analysis of stable C and H isotope composition is an important Philippines and Japan) only Δ13CH3D is available, thereby, not allowing
step in the study of methane origin, as widely demonstrated on biotic equilibrium or disequilibrium, and therefore the temperature, to be
(microbial and thermogenic) gas in surface biological ecosystems, assessed. The published data are shown in the Δ12CH2D2 vs Δ13CH3D
sedimentary gas-oil bearing systems (source rocks, reservoirs and seeps) diagram of Fig. 3.
and geothermal-volcanic manifestations. The fundamental empirical The data reveal that only two cases, Chimaera and Aqua de Ney,
δ13C–CH4 and δ2H–CH4 genetic diagrams of Schoell (1980) and Whiticar show isotopologue equilibrium or near-equilibrium (suggesting a for­
et al. (1986), and their updated versions of Milkov and Etiope (2018), mation temperature of around 130 ◦ C and 50 ◦ C, respectively; Young
are extensively used in gas geochemistry, and provide clues as well as et al., 2017; Blank et al., 2017), while the other cases fall above or below
debates about putative abiotic methane (Etiope and Sherwood Lollar, the equilibrium line (Fig. 3). The reasons of these deviations from
2013; Reeves and Fiebig, 2020), including methane from continental disequilibrium are elusive. Available data and models suggest that iso­
serpentinization systems (Etiope and Whiticar, 2019; Monnin et al., topologue disequilibrium, beyond being typical of relatively low tem­
2021). An updated version of δ13C–CH4 and δ2H–CH4 data from conti­ perature reactions, can be due to a series of genetic and post-genetic
nental serpentinization sites is shown in Fig. 2. The variability of processes: (a) kinetics, (b) mixing between abiotic endmembers or be­
δ13C–CH4 and δ2H–CH4 can be due to genetic (13C and 2H enrichment of tween abiotic and biotic (microbial or thermogenic) endmembers, (c)
the CH4 precursor, temperature and “maturation” of gas formation) or fractionation of isotopologues by molecular mass, (d) oxidation (e.g.,
postgenetic processes (desorption, oxidation, diffusion, mixing; Fig. S1 Anaerobic Oxidation of Methane, AOM), (e) combinatorial effects,
in Supplementary Material). It should be noted that this thematic and/or (f) tunneling (Wang et al., 2016; Young et al., 2017; Taenzer
approach principally relies on empirical fields for δ13C and δ2H, which et al., 2020; Young, 2020; Warr et al., 2021; Fig. S1). In particular, when
are continuously being redefined as new experimental and field data a reaction uses hydrogen from different sources, with a different 2H/H
comes to light. For example, Monnin et al. (2021) indicated that the free ratio, then a low Δ12CH2D2 results (combinatorial effect; Taenzer et al.,
(not dissolved) gas with δ2H–CH4 > − 200‰ is typical of dry gas seeps, 2020). Combinations of two or more of the processes (a) to (f) can lead to
with the exception of Gokdere (a spring in Turkey), while δ2H–CH4 < a wide range of Δ12CH2D2 vs Δ13CH3D values, including those typical of
− 200‰ is always from springs. We note that Aqua de Ney bubbling microbialgenesis. Therefore, it is not possible to define unambiguous
spring also has δ2H–CH4 > − 200‰ and gas dissolved in springs or genetic zonations in the Δ12CH2D2 vs Δ13CH3D diagram. It is worth
aquifers may have δ2H–CH4 > − 200‰ as well (e.g., Tablelands and noting that methane of Cabeço de Vide (Portugal) falls within the
CROMO). The variation of δ2H–CH4 seems, therefore, to be only microbialgenic zone, but methanogens were not detected in the water
partially linked to the interaction with the hyperalkaline water. (Tiago and Veríssimo, 2013). It is also worth noting that when only the
Notwithstanding the numerous postgenetic factors potentially affecting value of Δ13CH3D is measured, the assumption that methane formed at
the stable C and H isotope composition (Fig. S1), the 13C-enriched the relative “apparent” temperature is deceptive: for example, Suda
methane of continental ultramafic systems has a relatively homogenous et al. (2022) developed an abiotic CH4 origin model for the Happo site
isotopic genetic field, distinguished from thermogenic and microbial gas (Japan) assuming the apparent temperatures of about 200–300 ◦ C
of sedimentary basins. The Samail (Oman) gas, including the samples derivable from Δ13CH3D values of ~2.4 and 1.5. But similar values were
with positive δ13C–CH4 values, is however considered dominantly mi­ reported for Samail and Acquasanta gas (sites with lower geothermal
crobial by Nothaft et al. (2021). gradients than Happo), which are not in isotopologue equilibrium and
the real temperature cannot be defined. The case of the CH4 generated in
3.2. The CH4 clumped isotopes laboratory at 70 ◦ C (Young et al., 2017), whose apparent Δ13CH3D
related temperature would be ~150 ◦ C (Fig. 3), is illuminating.
The advent of multiply-substituted “clumped” CH4 isotopologues Clumped isotopes for methane of certain abiotic origin, as it is pro­
analyses (13CH3D and 12CH2D2, see review in Young, 2020) has allowed duced in laboratory, are available from three different experiments, with
additional genetic and post-genetic information to elucidate on tradi­ low and high temperature FTT synthesis and high temperature silane
tional “bulk” isotope analyses (δ13C and δ2H), specifically regarding the (Si5C12H36) decomposition (Young et al., 2017; Zhang et al., 2021). The
temperature of CH4 formation (Stolper et al., 2015; Young, 2020 and high temperature silane-derived gas is in isotopologue equilibrium,
references therein). CH4 clumped isotopes have helped to distinguish while the low temperature Sabatier reaction gas has a strongly negative
between microbial, thermogenic, and very high temperature abiotic Δ12CH2D2 value (Fig. 3). Interestingly, the 340 ◦ C FTT experiments of
(magmatic, hydrothermal) origins (e.g., Stolper et al., 2018; Wang et al., Zhang et al. (2021) showed methane initially with Δ13CH3D values from
2018; Giunta et al., 2019; Young, 2020). The assessment of the CH4 +1.3 to − 1.6‰ (anti-clumped) and Δ12CH2D2 = − 45‰, falling within
formation temperature is possible only when methane isotopologues are the microbialgenic zone in the plot of Fig. 3. Both clumped isotope
in thermodynamic equilibrium (Young et al., 2017). This generally values gradually increased, as for bond reordering, and finally
happens for methane formed at high temperatures, >150 ◦ C (thermo­ approached the thermodynamic equilibrium at 340 ◦ C in about 3 h. This
genic and magmatic/hydrothermal gas). Clumped isotope analyses experiment showed that during FTT synthesis the clumped isotopes can
revealed that microbial gas (from lab cultures and wetlands) is in iso­ evolve during the reaction, and that early methane produced and
topologue disequilibrium, with a specific range of combinations of low released from the system (reactants plus catalyst) is not in thermody­
to negative Δ13CH3D and negative Δ12CH2D2 values (Wang et al., 2015; namic (isotopologue) equilibrium. Obviously, there is the problem of
Young, 2020 also for details on the delta notation in per mil units). knowing if and how the early clumped-isotope signature, observed
Methane elsewhere with such a range of combinations (expressed as a rapidly after the start of the reaction, is observable in natural gas. We
genetic zonation in a Δ12CH2D2 - Δ13CH3D diagram; Fig. 3) is, therefore, don’t know whether fast evolution laboratory FTT products are appli­
considered very likely microbialgenic. Such a conclusion was applied to cable to natural FTT processes. The time-scale issue is a general problem
the methane sampled in some wells in the Samail ophiolite in Oman for any attempt to simulate nature with lab experiments. In this respect,
(Nothaft et al., 2021), an area where the gas, sampled in surface seeps we can consider processes associated with thermogenic methane pro­
and springs, was formerly considered mainly abiotic (Fritz et al., 1992; duced in and released from shales and coal, and apply them here. There
Vacquand et al., 2018), and which will be discussed in more detail in are three processes inducing isotopic evolution of gas observable from
Section 5. shales and coal: (a) maturation; (b) desorption; (c) diffusion.

5
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

(a) Methane produced by catagenesis follows a 13C-increase evolu­ km wide, are in the Voltri ophiolite (Italy), Samail ophiolite (Oman),
tion (maturation), so that the first methane molecules that are Ronda peridotite massif (Spain), Dinaride ophiolite (Bosnia-Herzego­
produced (early mature) and liberated out of the shale/coal are vina) and Kizildag ophiolite (Estern Turkey; Table S1 and references
not in thermodynamic equilibrium and are more 13C depleted; therein). These systems often reveal the existence of crustal degassing
methane produced later, over longer time scales will be more 13C processes, which should be considered to interpret local data. A common
enriched. The effects of this on clumped-isotope composition are (or dominant) origin of methane may exist in all gas occurrences of a
discussed by Xia and Gao (2019). given regional system. Examples of regional contextualizations are in
(b) Once formed in shales or coal, much of methane is adsorbed on Etiope et al. (2016), D’Alessandro et al. (2018) and Vacquand et al.
the surface of minerals and organic matter. When depressuriza­ (2018).
tion begins (for geological reasons, uplift, changes of lithostatic For the interpretation of local data, a global perspective on gas in
pressure, etc.) methane enriched in 12C desorbs more rapidly than continental serpentinization systems is also fundamental. An example of
13
C-enriched methane. The residual methane, liberated in later this type of consideration was explored in Monnin et al. (2021), where
stages, is more 13C-enriched. This process is also a cause of they compared molecular and isotopic gas composition from different
isotope trend reversal, because residual methane becomes continental serpentinization sites. Differences between methane dis­
enriched in 13C more rapidly than ethane (Milkov et al., 2020). solved in water and in dry gas seeps and specific relationships between
(c) Free gas in shale/coal migrates towards other rocks by diffusion- methane and other gases (H2 and N2) were identified. A gas origin
advection (generally driven by decompression), and the diffusive sequence model was proposed, whereby, H2 is first produced in water by
movement induces isotopic fractionation (i.e., the first methane serpentinization, then it reacts in gas-phase with CO2 (Sabatier reaction)
diffusing out is more 12C-enriched). producing CH4. The methane can subsequently migrate and dissolve in
water where microbial (methanotrophic and methanogenic) activity
In natural FTT synthesis it is plausible to have similar processes, takes place. This coincides with the CO2 hydrogenation – reservoir
whereby, the degree of FTT reaction can be seen as a “maturation” (the model of Etiope and Whiticar (2019).
first liberated gas is early mature and not in isotopologue equilibrium); Some holistic questions are suggested here only for reflection pur­
desorption is actually a step foreseen in FTT reactions (desorption from poses. Can the three theories of gas origin (i.e., microbialgenesis, abiotic
the catalyst); what Zhang et al. (2021) observed is likely a first liberation from inclusions, abiotic from CO2 hydrogenation-reservoir) be valid
(desorption) of methane, which is in isotopologue disequilibrium. Zhang systematically, everywhere in all 20 countries where methane is
et al. (2021) also found an increase of δ13C over time, similar to the low observed to stem from ultramafic rocks (Table S1, and the map in Etiope
temperature Sabatier experiments (Etiope and Ionescu, 2015). Stolper and Whiticar, 2019), or are they complementary, each of them valid
et al. (2018) mentioned the possibility that desorption from shales could only in some cases and not in others? For example, can Chimaera gas be
be expressed as kinetic isotope effects influencing methane dominantly abiotic, while Oman gas is microbial and Philippines gas is
clumped-isotope composition. Diffusion can also induce mainly from inclusions? Should we look for a unique (dominant) gas
clumped-isotope fractionation (Fig. S1). origin, common to all cases, or all three origins may occur simulta­
Basically, we cannot exclude the possibility that, also in natural neously, in variable proportions, in all sites, so that the gas is always
geological conditions, early abiotic gas generated in a CO2 hydrogena­ mixed? Mixtures of the three origins at the same site are theoretically
tion reaction may have low clumped-isotope values, and that residual possible because all ophiolites or peridotite massifs can have all items
(late-desorption) and/or mature abiotic gas has higher values. Accord­ needed for the three origins: present-day serpentinization (= H2 and
ingly, residual and/or mature abiotic gas would be in isotopologue energy sources for methanogens), peridotites with fluid inclusions,
equilibrium, if formed at relatively high temperature; early abiotic gas metal (catalyst)-rich rocks and fracture networks for migration of gas
would be always in disequilibrium and could have Δ13CH3D and (including inputs of CO2 from external sources). Nevertheless, it is
Δ12CH2D2 values similar to those of microbial gas. plausible that one genetic process dominates at the global scale: this
could be preliminarily identified by weighing the pros and cons deter­
4. Holistic contextualization of local observations mined by available data (Fig. 4), and considering one fundamental
factor, often neglected in gas-geochemical studies: how to explain the
The interpretation of CH4 origin at a given site cannot be based solely intense and long-lasting methane flow observed at the surface. This is
on what is observed in that site. If a gas seep or gas-rich aquifer is part of discussed in the next sections.
a wider regional gas system and similar systems exist at global scale,
then the whole regional-global system must be considered to understand 4.2. The clumped-isotope perspective
the local processes. Geochemical data should include gas species that,
even if they are not directly related to methane origin (e.g., N2, noble If clumped-isotopes are used to unravel the gas origin, we note that
gases), may reveal crustal or mantle degassing processes. Geochemical methane in three serpentinization sites, Acquasanta (Italy), Aqua de Ney
data should then be integrated with geological, structural and petro­ (California) and Chimaera (Turkey), do not fall within the micro­
graphic factors, e.g., the relationships between gas discharges and faults, bialgenic zone of Fig. 3, and, therefore, should not be considered
distribution of ultramafic rocks (especially those with metals that can be dominantly microbial (unless other processes are invoked to demon­
potential catalysts of FTT reactions), C bearing rocks and sedimentary strate that also microbial gas can have non-microbialgenic clumped-
formations. Not least, the interpretation of gas origin should be isotope signatures). So, based on the current clumped isotope interpre­
compatible with quantitative aspects, such as the gas flow and emission tation (e.g., Young, 2020; Nothaft et al., 2021), not all sites are domi­
on the surface. A holistic approach may offer some answers, as discussed nated by microbial gas. We note also, as mentioned above, that Cabeço
below. This approach is then used to critically review the case of Samail de Vide (Portugal) falls within the microbialgenic zone, but it lacks
gas in Oman, as discussed in Section 5. methanogens (and has 13C-enriched CH4 and C2+ alkanes; Tiago and
Veríssimo, 2013; Etiope et al., 2013b).
4.1. Regional and global contextualization of local data Methane from fluid inclusions (high temperature hydrothermal or
post-magmatic origin) should be in isotopologue equilibrium and indi­
In several cases, the occurrence of gas in a specific site, seep, spring cate a relatively high temperature, like the data from submarine hy­
or aquifer, is the local expression of a wide, regional gas seepage system. drothermal fields (Fig. 3). Only two sites, Chimaera and Aqua de Ney,
Examples of regional seepage systems in ultramafic rocks, with multiple are in or near isotopologue equilibrium, but at temperatures lower than
gas seeps or springs distributed over areas hundreds to thousands square 150 ◦ C. The Oman and Portugal sites, being below the equilibrium line,

6
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

are not likely dominantly from inclusions. Based on clumped isotopes, 4.4. Fluid inclusions as a source of surface gas seeps
no site gas would dominantly come from inclusions; unless post-genetic
processes modify the clumped-isotope composition. The presence and concentration of methane in fluid inclusions has
Methane from CO2 hydrogenation would have a wider clumped- created debate surrounding its origin as well as how it may supply a
isotope signature range, depending on the disequilibrium-equilibrium continual and pressurized gas source observed at the surface in the form
evolution during the catalyzed reaction, involving desorption and of dry gas seeps and/or associated with springs (Etiope and Whiticar,
diffusion fractionations, as discussed above. Accordingly, the clumped- 2019; Klein et al., 2019). Olivine hydrolysis experiments, direct mea­
isotope composition would not be a single diagnostic of abiotic CO2 surements of methane (and other gases) in olivine fluid inclusions, and
hydrogenation. The data related to the several continental serpentini­ isotope analyses from field locations provide the basis/start to deter­
zation sites are actually sparse in the Δ13CH3D vs Δ12CH2D2 plot (Fig. 3). mine how olivine can directly contribute methane in these systems. The
complexity of this issue can be highlighted by the case history of the San
4.3. The quantitative factor: gas flow dynamics constraints Carlos olivines (Arizona, USA), which were studied to assess whether
methane in inclusions can represent a background source in laboratory
A key factor that should not be neglected is the amount of methane FTT experiments. In these olivines, McCollom (2016) measured a wide
emission at the surface. Many continental serpentinization seeps or range of methane concentrations, by a factor of 49. But Pasteris and
springs release significant amounts of methane, such as the Chimaera Wanamaker (1988) and Smith et al. (2017) did not detect any methane.
seep in Turkey (several hundreds of tonnes per year; Etiope and Schoell, The presence of CH4 in microfractures and veins in igneous rocks, rather
2014), the Archani-Ekkara springs in Greece or the Acquasanta seepage than in inclusions, has been observed in several cases (e.g., Potter and
system in Italy (hundreds of kg per year; Etiope et al., 2013a; Etiope and Konnerup Madsen, 2003; Nivin, 2016), including ultramafic rocks
Whiticar, 2019). Methane fluxes were not measured at the >60 (Etiope et al., 2018). Therefore, modeling how and to what extent ol­
springs/seeps in Oman, but the water flow rates, active bubbling, dry gas ivines can contribute methane solely from fluid inclusions is difficult,
seeps and the wide area of seepage suggest that the overall emission is especially when attempting to scale up.
relevant. Other sites with significant transport of methane to the surface, Expanding on other studies, the abundance of methane determined
either through multiple springs, bubble plumes or dry gas vents, are by crushing whole rocks (bulk analysis), as performed by Grozeva et al.
documented in Eastern Turkey (D’Alessandro et al., 2018), Southern (2020), does not represent exclusively gas from inclusions. To be sure
Spain (Etiope et al., 2016), Philippines (Los Fuegos Eternos; Vacquand that in crushing analyses gas derives solely from inclusions, it would be
et al., 2018) and Indonesia (Tanjung Api “eternal flames”; Subroto et al., necessary to completely isolate olivine crystals, as done in Salvi and
2004; Fig. S4). Can these continuous and long-lasting emissions be Williams-Jones (1997). In the context of fluid inclusions, what is true for
supported by the three invoked methane sources individually (micro­ one olivine is not the same for olivine sampled from the same location,
bial, abiotic inclusions, CO2 hydrogenation-reservoir model)? Visible let alone olivine from different locations.
gas emissions, either as bubbling in water or as dry gas seeps, are clearly There is, then, a debate on whether fluid inclusions can support the
advective, driven by pressure gradients. This implies that the gas source large amount of methane released over 2000 years from the Chimaera
is pressurized, from accumulation, or that the gas generation process seep in Turkey (Etiope and Whiticar, 2019; Klein et al., 2019); the dis­
(the “kitchen”) is very prolific, continuous and fast enough to maintain cussion was around approximate calculations (Etiope and Whiticar,
the surface output. We think this second hypothesis is not realistic, as 2019) based on extreme assumptions and values of parameters
gas production rates in geology are typically very slow. The advantage of (maximum volume of fluid inclusions in igneous rocks, 0.01% of rock, as
having a reservoir, like in the CO2 hydrogenation-reservoir model, is indicated by Roedder (1972); 100% of CH4 in each inclusion; very
that the gas production must not necessarily be continuous and intense; conservative amount of methane, 0.5 km3, produced over only 2000
the gas can progressively form over geological time-scale, with rela­ years) considered just to derive the very minimum theoretical volume,
tively low production rates (as estimated in Etiope and Whiticar, 2019), in terms of order of magnitude, of rock containing inclusions. Using a
and gradually migrate towards permeable rocks, exactly as in a petro­ range of methane partial pressure observed in inclusions, Klein et al.
leum system. The microbial and fluid inclusion theories need to be (2019) derived a range of 1383 to 10 km3 of rock, and considered that
completed with a satisfactory gas flow model. In particular, the aquifer fluid inclusions cannot be excluded as a source of the Chimaera gas.
microbial theory should explain how methane can accumulate (with Details of the calculations and related debate are reported in the Sup­
pressure) above the aquifer (Fig. 4). The microbial model shall, then, plementary Material. Whether or not the range 101–103 km3 of rock is
explain where (at what depth and in which rock) methanogens operate realistic (considering the need to convey the gas from this rock volume
and how can they feed, also in the spatial dimension, the multiple gas to a focused area on the surface, continuously and for very long time, as
discharges with advective flows (via gas accumulations or through fast discussed in Section 4.4), we reiterate that the derived numbers refer to
gas generation?). The C-feedstock of methanogens should also be extreme values, not representative of the average rock conditions.
enough to match the final CH4 output. A picture of the carbon cycle, Methane concentration is not 100% in all inclusions, the volume of fluid
examining availability of formate, acetate, CO and atmospheric CO2, inclusions in igneous rock is often <0.01% of rock and the amount of
should be developed. Doubts exist, in fact, on sufficient availability of methane released by the Chimaera seep since its origin (which certainly
organic acids and CO compared to the amounts of CH4 released (Etiope predates 2000 years) is much higher than 0.5 km3. Using average values
and Whiticar, 2019). The fluid inclusion model shall describe how and of the several parameters, volumes of peridotite rocks at least in the
why methane is leached from fluid inclusions, the volumes of rocks with order of 104 km3 of rock can be derived (Supplementary Material, S1).
the gas-rich inclusions that should be involved and how, once a given There is a need for more realistic values and modeling that still needs to
volume of inclusions is evacuated, new inclusions in other volumes of be completed.
rocks (laterally or vertically distributed) are involved, always producing
the relevant and continuous gas flows observed at the surface. Basically, 5. The case of methane in the Samail ophiolite
how would fluid inclusions continuously rupture/leak on such a time
and geographic scale? This fluid dynamics perspective is essential in An interesting and complex case of multiple interpretations for
order for the multiple gas origin hypotheses to be both geologically and methane origin in ultramafic rock settings is the one of the aquifers and
physically realistic. springs in the Samail (or Semail) ophiolite in Oman. This region hosts
numerous sites, probably more than 60, with hyperalkaline springs and
dry gas seeps where methane is documented together with H2 and N2
(Neal and Stanger, 1983; Fritz et al., 1992; Sano et al., 1993; Vacquand

7
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

et al., 2018 and references therein). Vacquand et al. (2018) showed that of fluid inclusions supporting the abundance of alkanes detected in the
a complex gas-water seepage system exists along a belt of >350 km wells, but refer to Miura et al. (2011), who reported Raman analysis of
comprising several massifs of the Samail ophiolite, and that the gas has only one harzburgite sample from another massif, about 280 km north of
variable CH4, H2 and N2 compositions. In all cases, a dominant abiotic the area (the Wadi Tayin Massif) studied by Nothaft et al. (2021). Miura
CH4 component was suggested based on multiple gas-geochemical data, et al. (2011) reported that the inclusions are secondary, referring to gas
in agreement with earlier studies (Fritz et al., 1992; Sano et al., 1993). trapped in healed fractures, and that most of the inclusions are gas-free.
The recent study by Nothaft et al. (2021) focused on fluids intercepted Using this reference (one sample, with rare CH4-inclusions, far from the
by a series of wells drilled (max depth of 400 m) in the framework of the studied area) to support the hypothesis that methane in the Samail
Oman Drlling Project, in the southern Wadi Tayin Massif, near Ibra. ophiolite comes from inclusions in local peridotites is, at least,
Based on microbiological (16S rRNA Gene Sequencing) and fluid premature.
geochemical analyses including CH4 clumped isotopes, Nothaft et al. There are no studies on the methane-inclusion mass balance in pe­
(2021) claimed that methane in the Samail ophiolite is mainly micro­ ridotites. The capability of the inclusions to support long-lasting and
bial, with a minor abiotic component derived from fluid inclusions in the large gas emissions on the surface, as observed in some places (e.g., the
peridotite. The microbial origin was suggested on the basis of: (a) the Chimaera burning seep in Turkey) is the subject of a specific debate
presence of methanogens in the water intercepted by the investigated (Etiope and Whiticar, 2019). A detailed discussion on this aspect is re­
wells (water pumped from depths mostly <100 m), and models pointing ported in Section 4.4 and Supplementary Material. A critical question,
to energetically favourable formatotrophic methanogenesis. Metha­ discussed in Section 4.3, concerns the gas flow mechanics: what is the
nogens can produce 13C-enriched CH4 in C-limited conditions (perido­ mechanism of methane leakage from inclusions that involves progres­
tite aquifer); (b) low values of CH4 clumped isotopes Δ12CH2D2 and sively different portions of the peridotite massif, after that the inclusions
Δ13CH3D (considered to be characteristic of microbialgenesis, as dis­ from a certain volume of rock are evacuated, and assuring continuous
cussed above). advective gas flows driven by intense pressure gradients? A
The subordinate fluid inclusion origin was suggested by: (c) the geological-fluid dynamic model, including mass-balance, should be
occurrence of 13C-enriched methane, ethane and propane, similar to developed to answer the question. For the Samail ophiolite seepage belt
those found in seafloor hydrothermal vents where fluid inclusions are (>350 km long), the continuous release of methane in at least 60
considered a dominant source of alkanes (e.g., Grozeva et al., 2020); (d) different places (whether from geological or historic times is unknown)
literature reference of some, but not all applicable, CH4-bearing in­ would require an incredibly huge volume of rock with high concentra­
clusions found in the Samail peridotite (Miura et al., 2011). Nothaft et al. tions of CH4-rich inclusions.
(2021) did not include the abiotic CO2-hydrogenation-reservoir model
(where chromitites can be gas source rocks), claiming that: (e) there is 5.2.2. The meaning of the C1/(C2+C3) ratio
not enough free H2 or CO2 for FTT catalysis; (f) there are no chromitites The gaseous alkanes heavier than methane (i.e., ethane, propane,
(catalyst-rich rocks) in the area, neither in rocks intercepted by the and occasionally butane) have been widely reported in continental
investigated wells or in surrounding outcrops. Points (b) to (f) are serpentinization seeps and springs. Their presence was documented in
examined in detail in the following sections. Local data are contextu­ all sites where clumped-isotopes were also measured (Fig. 3). The C1/
alized considering the wider Samail gas seepage system and the valuable (C2+C3) ratio (methane/(ethane + propane), also called “Bernard
addition of measuring the radiocarbon (14C) content of CH4 is discussed. ratio”) is frequently used to support the δ13C vs δ2H diagram to better
understand the gas origin. An updated version of the genetic zonations
5.1. Low values of CH4 clumped isotopes Δ12CH2D2 and Δ13CH3D are not in the δ13C vs C1/(C2+C3) diagram, based on a wide biotic and abiotic
exclusive of microbialgenesis dataset, is reported in Milkov and Etiope (2018). While this diagram can
be mostly useful to distinguish (with some caution) between microbial
Clumped isotopes of methane sampled in three wells in the Samail and thermogenic gas, it has no value for gas in igneous rocks,
ophiolite have a Δ13CH3D versus Δ12CH2D2 combination typical of hydrothermal-magmatic systems and serpentinized ultramafic rock
microbialgenesis (Nothaft et al., 2021). This is a major point used by systems (putative abiotic gas), since this gas has a wide range of
Nothaft et al. (2021) to support their microbial origin hypothesis. This C1/(C2+C3) values (Milkov and Etiope, 2018). In addition, it is known
point loses strength when a more rigorous analysis of the numerous that the C1/(C2+C3) ratio is an ambiguous genetic proxy because ethane
factors influencing the clumped isotopic composition and the recent and propane concentrations can decrease during gas migration (mo­
laboratory FTT data, discussed above, are considered. Mixing, frac­ lecular fractionation), due to differential adsorption and solubility (e.g.,
tionation of isotopologues by molecular mass, desorption, oxidation, Deville et al., 2003; Etiope et al., 2009). For example, gas in surface
kinetics and tunneling (Young, 2020), complicate the interpretation of seeps often has less ethane and propane, relative to methane, than in
clumped isotopes and can lead to signatures similar to those of micro­ reservoirs, which in turn have less ethane-propane compared to the
bialgenesis (Fig. S1). The FTT experiments of Young et al. (2017) and source rocks (Milkov and Etiope, 2018). It seems that the C1/(C2+C3)
Zhang et al. (2021), and the case of Cabeço de Vide mentioned above ratio on the surface depends on the intensity, velocity of the gas flow
(where methanogens were not detected), actually showed that abiotic from the reservoir or source rock (the higher the velocity, the lower the
gas can have Δ13CH3D and Δ12CH2D2 signature similar to microbial gas residence time in the subsurface, the lower the molecular factionation).
(Fig. 3). The Samail ophiolite CH4 requires additional clumped isotope The same can happen for abiotic gas. Furthermore, abiotic gas produced
analyses, especially of samples taken in surface discharges, preferably by FTT synthesis can have a wide range of C1/(C2+C3) ratios, since the
the dry gas seeps (as those studied in Vacquand et al., 2018), which are amount of several alkanes depends on the degree of methane polymer­
not considered in Nothaft et al. (2021). ization, which is function of temperature and time (Sherwood Lollar
et al., 2008). Therefore, the C1/(C2+C3) ratio of abiotic gas does not
5.2. The fluid inclusion hypothesis have a specific signature.
In the gas observed in the Samail ophiolite well NSHQ14, there are
5.2.1. Existence and abundance of fluid inclusions significant amounts of ethane and propane, and this led Nothaft et al.
In the Samail ophiolite, an additional gas origin from fluid inclusions (2021) to consider an additional non-microbial gas component. It is
was hypothesized by Nothaft et al. (2021) to explain the minor presence worth noting that Samail gas also has heavier alkanes, butane to hexane
of 13C-enriched ethane and propane in the observed wells. Although the (Neal and Stanger, 1983; Nothaft et al., 2021), which would turn the
authors write that this “is evidenced by the presence of CH4-bearing fluid attention towards a possible sedimentary, thermogenic origin. Nothaft
inclusions in the Samail Ophiolite,” their work did not report any analysis et al. (2021) reported a C1/(C2+C3) ratio from 1240 to 881, and claimed

8
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

that these values are not representative of low temperature abiotic tectonite basal ultramafic member. Etiope et al. (2018) found that
origin, because they are more similar to those in fluid samples and rocks among all rocks composing four different ophiolites in Greece, only
from other ophiolites and sediment-poor seafloor hydrothermal vents, chromitites host relevant amounts of methane, hydrogen and C2+
and are much higher than those reported for low temperature abiotic gas heavier alkanes. A similar finding was reported by De Melo Portella et al.
observed in Precambrian shields (Kidd Creek, Canada). As explained (2019) for chromitites in a Archean-Paleoproterozoic greenstone belt in
above, however, this approach to decipher the gas origin by comparing Brazil. Raman analyses showed that methane in chromitites is adsorbed
C1/(C2+C3) with significantly different abiotic systems, without along microfractures in samples enriched with chromium and ruthe­
considering the context, risks misinterpretation. nium, two powerful catalysts of Sabatier reaction. Hence, the hypothesis
In the Table 4 presented in Nothaft et al. (2021), it seems that the that chromitites may act as a gas source rock (Etiope et al., 2018; Etiope
C1/(C2+C3) ratio was calculated using dissolved gas concentrations and Whiticar, 2019).
values (which depend on the solubility of the gas), which cannot be Concerning the Samail ophiolite, Nothaft et al. (2021) claimed that
compared to the ratios derived from free gas phases, as completed for the chromitites cannot be the source of the methane observed in the wells,
Kidd Creek gas (and as generally done for the Bernard diagram). Un­ because massive chromitites do not exist in the region as they are not
fortunately, the original composition of gas extracted from water was reported in the cores or drill cuttings from any well in the area, or in
not reported. The fluid inclusion hypothesis shall then be better justified outcrops in the surrounding area. This is not correct. The authors missed
by explaining how butane to hexane alkanes form in the inclusions. the studies with chromitite mapping and PGE (Platinum Group Ele­
ments) analyses of Page et al. (1982) and Rollinson and Adetunji (2015).
5.2.3. The stable C isotopic composition of ethane and propane Page et al. (1982), in particular, report at least eight chromitite outcrops
13
C-enrichment in ethane and propane is generally an indication of (where PGEs were observed) in the area investigated by Nothaft et al.
the high maturity of thermogenic gas. It may also be an effect of post- (2021); one outcrop is 5 km far from the NSHQ14 well (Fig. S2).
genetic processes (oxidation, isotopic fractonation by diffusion). The Chromitites, containing PGE, in the Ibra area correspond to the basal
latter may apply to abiotic gas. For the Samail ophiolite gas, Nothaft ultramafic member (predominantly harzburgite and dunite tectonite),
et al. (2021) report a stable C isotopic composition of ethane and pro­ which develops underground in other portions of the Wadi Tayin Massif
pane of only one sample (NSHQ14 – 2017; δ13C2: +6‰, δ13C3: +3.3‰). (Page et al., 1982; Rollinson and Adetunji, 2015).
Combined with the non-detection of C2–C3 oxidising bacteria in the Moreoever, Nothaft et al. (2021) assumed that there is no free H2
NSHQ14 well water, the 13C-enrichment was considered to reflect the available for CO2 hydrogenation, as the subsurface is water-saturated.
formation conditions. As the isotopic values are similar to those However, underground rocks are not always water-saturated. Any
observed in fluid inclusions in rocks from the oceanic lithosphere, it was aquifer does have a bottom level. Examples of water-free rock systems
assumed that the gas cannot be derived from low temperature CO2 hy­ include gas-filled fractures and pores observed: a) in petroleum fields, b)
drogenation. This is only one of the several solutions of the problem. in 3D microscopy of immiscible displacement in rocks (Porter et al.,
Other processes leading to 13C-enriched C2 and C3 cannot be excluded a 2015) and c) in gas migration studies of gas columns, slugs and bubbles
priori: along fracture networks (e.g., Malmqvist and Kristiansson, 1984). In
stagnant, serpentinized systems, Sleep et al. (2004) postulated that
(a) If alkanes are not produced in the water (as Nothaft et al., 2021 under certain conditions H2 fugacities can exceed saturation. H2(gas)
assumed considering the fluid inclusion origin), they could be bubbles and gas-filled fractures could exist under such conditions.
oxidised before dissolving into the investigated water, for Nothaft et al. (2021) do not mention that ~5 km from the well CM2A,
example within deeper pore/fracture water. there is a H2-rich bubbling spring, named Lauriers Roses, or Nasif
(b) C2–C3 may be derived from polymerization of 13C-enriched CH4 (Fig. S2). This site releases free gas with 61% of H2, 23% of N2 and 15%
(Sherwood Lollar et al., 2008). 13C-enrichment in CH4 can be, in of CH4 with δ13C–CH4: +7.9‰ (Vacquand et al., 2018). Not considering
theory, due to methane oxidation or the formation from a this site, as well as the numerous springs and bubbling seeps that are
13
C-rich precursor. Such a precursor could be the CO2 derived by widespread in the Samail ophiolite, and which have variable N2–H2–CH4
thermal decarbonation (Wycherley et al., 1999) from the compositions, makes any interpretation uncontextualized.
13
C-enriched Permian carbonates, such as the Hawasina, The availability of CO2 should also be examined in the regional
Sumeini, or Saih Hatat formations occurring below and in lateral context. While CO2 does not (obviously) exist in the hyperalkaline fluids
tectonic contact with Samail ophiolite nappe (Hopson et al., intercepted by the wells, it can be pervasive in the metasedimentary
1981). rocks below the ophiolitic nappe, and migrate upwards through the
(c) 13C-enriched C2–C3 (as well as CH4) may represent residual gas numerous fault and fracture systems of the region (Vacquand et al.,
isotopically fractionated due to late desorption and diffusion in 2018). These fracture systems can be the loci of H2 and CO2 mixing. It is
low permeability rocks (as discussed above with reference to the worth noting that, as mentioned in Section 5.2.3, 13C-enriched CO2 may
evolution of clumped-isotope composition). 12C-enriched gas be derived by thermal decarbonation from the 13C-enriched Permian
desorbs and diffuses faster than 13C-enriched gas; residual gas is carbonates, such as the Hawasina, Sumeini, or Saih Hatat formations
then 13C-enriched. This phenomenon is typically observed in occurring below and in lateral tectonic contact with the Samail ophiolite
sedimentary gas source rocks, but it can occur in any low nappe. Hydrogenation of 13C-enriched CO2 may then lead to
13
permeability system. C-enriched methane.

Accordingly, 13C-enriched ethane and propane in the Samail ophio­ 5.4. Gas outside the observed aquifer
lite NSHQ14 well may not necessarily come from fluid inclusions, whose
abundance in the area investigated was not assessed, as discussed in The assumption behind the hypothesis of a microbial origin is that
Section 5.2.1. CH4 is mostly produced within the peridotite aquifer. A wider look at the
regional context in Oman reveals that the gas seepage system extends
5.3. Availability of chromitites and free reactants for catalyzed CO2 beyond the local peridotite aquifer of the Wadi Tayin Massif (Fig. S3).
hydrogenation For example, N2 is another missing item in the analyses of the Samail
ophiolite wells in Nothaft et al. (2021). N2-rich springs are abundant in
CO2 hydrogenation needs metal catalysts, therefore, methane should the region (e.g., Boulart et al., 2013; Vacquand et al., 2018); in these
be produced in a metal-rich rock. In ophiolites, potential metal catalysts springs N2 is mostly of crustal, sedimentary origin, with a minor atmo­
are particularly abundant in chromitites, typically occurring in the spheric component, as evidenced by δ15N isotopes (Vacquand et al.,

9
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Fig. 5. Flow chart showing how the several techniques and analyses may guide the unravelling of methane origin.

2018). Basically, N2 comes from sediments below the ophiolitic units, reservoir model would be compatible, and a microbial origin could be
from clay dewatering or destabilization of ammonium bearing clays. The considered only if methanogens use a 14C-free substrate. On continents,
existence of crustal degassing is then suggested by relatively high con­ potential C feedstock for methanogens are carbonate veins, typically
tents of 4He (Vacquand et al., 2018). This testifies that the gas seepage pervading the peridotites (Suzuki et al., 2014), atmospheric CO2 and the
system crosses the whole Samail ophiolitic nappe and includes deep products of present-day serpentinization, formate (HCOO− ), acetate
sedimentary inputs. With respect to the CH4-rich serpentinized ultra­ (CH3COO− ) and CO (Miller et al., 2018; Nothaft et al., 2021). In the
mafic complex of Hakuba-Happo in Japan, a similar approach Samail ophiolite, both carbonate dissolved in groundwater and car­
combining noble gases and carbon isotopes was applied to identify bonate veins have measurable 14C content (Abdalla et al., 2018; Nothaft
subsurface fluid components and ultimately identify crustal degassing as et al., 2021 and references therein). If methane is 14C-free, carbon in
the dominant carbon source (Suda et al., 2022). Can we exclude that the water and carbonate veins cannot represent a substrate for metha­
sedimentary input below the Samail ophiolite, in addition to N2, may nogens; then formate, acetate and CO feeding the microbes should be
also bear CO2 and hydrocarbons into the overlying ultramafic rocks? “ancient” and 14C-free, thereby, contributing to the dissolved carbon in
14
C-free aquifers, which have not been identified so far (Abdalla et al.,
14 2018). Analysis of 14C–CH4 is a high priority target as it may help clarify
5.5. Is Samail ophiolite methane C-free?
the origin of methane in many cases.
A key and missing parameter, not investigated so far in the Samail
6. Concluding remarks
gas, is the amount of radiocarbon (14C) in CH4. Measurable amounts of
14
C indicate that methane is formed from a “modern” C precursor, while
This review is meant to stress the importance of examining gas in
absence of 14C (expressed as 0 percent of modern carbon) means that the
ultramafic rock systems with a holistic approach, considering a wider
CH4 carbon is older than ~50,000 years (fossil C). 14C–CH4 analyses in
geological-geochemical (regional to global) context, with a vision that
several CH4-discharge serpentinization sites, Zambales (Philippines),
goes beyond the local observation. Specifically, the attribution of an
Ronda (Spain), Chimaera (Turkey), Acquasanta (Voltri, Italy), Happo
origin to a certain gas observed in a given system cannot be solely based
(Japan) and in the submarine Lost City (Abrajano et al., 1990; Pros­
on the data observed in that system, but shall consider what is observed
kurowski et al., 2008; Etiope and Schoell, 2014; Etiope et al., 2016;
in other geologically and geochemically similar places and, in particular,
Etiope and Whiticar, 2019; Suda et al., 2022), always revealed that
the elevated migration potential of gas, because the genetic history of a
methane is 14C-free. In contrast, the carbon precipitated or dissolved in
gas is often untied from the environment where it is sampled. The three
the related hyperalkaline springs contains substantial 14C, dating the
main theories of methane’s origin in continental ultramafic rock systems
water circulation of several hundred to a few thousand years (Cipolli
(i.e., abiotic CO2-hydration-reservoir model, fluid inclusion model and
et al., 2004; Marques et al., 2018; Etiope et al., 2016; Suda et al., 2022).
microbial-aquifer model) have their pros and cons. Some geochemical
Therefore, methane cannot be sourced from any of the 14C-bearing
data, taken individually, can be compatible with all three models, but
carbon dissolved in hyperalkaline waters.
some models are clearly incompatible with other data (Fig. 4). In
In the Samail ophiolite, 14C–CH4 has yet to be measured. If methane
particular:
has modern C, with measurable 14C, the deep origin models (i.e., fluid
inclusion and CO2 hydrogenation-reservoir models) would not be suit­
- methane clumped-isotopes are ambiguous genetic proxies for abiotic
able, as they consider a geological C precursor. The microbial origin
gas produced at moderate temperature: this gas is typically in iso­
hypothesis, with methanogens using modern C substrates, would be
topologue disequilibrium, and there is strong possibility that FTT
compatible. If methane is 14C-free, like in other continental serpentini­
reactions in natural systems give low/negative values of Δ13CH3D vs
zation sites, then both the fluid inclusion and CO2 hydrogenation-

10
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Δ12CH2D2 (as observed in laboratory) undifferentiated from References


microbialgenesis.
- Radiocarbon analysis of methane is a fundamental step in the Abdalla, O.A., Al-Hosni, T., Al-Rawahi, A., Kacimov, A., Clark, I., 2018. Coupling isotopic
and piezometric data to infer groundwater recharge mechanisms in arid areas:
interpretation of gas origin, as it precludes microbial origin from example of Samail Catchment, Oman. Hydrogeol. J. 26, 2561–2573.
groundwater that contain 14C. Radiocarbon-CH4 analyses are Abrajano, T.A., Sturchio, N.C., Kennedy, B.M., Lyon, G.L., Muehlenbachs, K., Bohlke, J.
required in more sites, and models and analyses that may verify K., 1990. Geochemistry of reduced gas related to serpentinization of the Zambales
ophiolite, Philippines. Appl. Geochem. 5, 625–630.
whether methanogenes may use a14C-free substratum need to be Baraj, E., Vagaský, S., Hlinčík, T., Ciahotný, K., Tekač, V., 2016. Reaction mechanisms of
developed. carbon dioxide methanation. Chem. Pap. 70, 395–403. https://doi.org/10.1515/
- The existence of gaseous alkanes heavier than methane (i.e., ethane, chempap-2015-0216.
Barbier, S., Huang, F., Andreani, M., Tao, R., Hao, J., Eleish, A., Prabhu, A., Minhas, O.,
propane and butane) must be interpreted considering that these Fontaine, K., Fox, P., Daniel, I., 2020. A review of H2, CH4, and hydrocarbon
gases can be affected by molecular fractionation processes during formation in experimental serpentinization using network analysis. Front. Earth Sci.
migration, including differential solubility. Additionally, a wide 8, 209. https://doi.org/10.3389/feart.2020.00209.
Blank, J.G., Etiope, G., Stamenkovic, V., Rowe, A.R., Kohl, I., Li, S., Young, E.D., 2017.
range of alkane composition can be produced by FTT processes,
Methane at the Aqua de Ney hyperalkaline spring (N. California, USA), a site of
depending on temperatures, catalysts and time evolution (e.g., by active serpentinization. In: Astrobiol. Sci. Conf. 2017, Abstract N. 3608, April 24–28,
polymerization); therefore, their relative composition (specifically 2017, Mesa, Arizona.
the C1/(C2+C3) ratio) observed in shallow aquifers or far from their Boulart, C., Chavagnac, V., Minnin, C., Delacourt, A., Ceuleneer, G., Hoareau, G., 2013.
Difference in gas venting from ultramafic-hosted warm springs: the example of
putative source rocks, cannot be used as a genetic index. Oman and Voltri ophiolites. Ofioliti 38, 143–156.
Brazelton, W.J., Thornton, C.N., Hyer, A., Twing, K.I., Longino, A.A., Lang, S.Q., et al.,
Fig. 5 is a flow chart of holistic interpretation of data, suggesting how 2017. Metagenomic identification of active methanogens and methanotrophs in
serpentinite springs of the Voltri Massif, Italy. PeerJ 5, e2945. https://doi.org/
the several techniques and analyses may guide the unravelling of 10.7717/peerj.2945.
methane origin. The major constraint in any interpretation of methane Brovarone, A.V., Martinez, I., Elmaleh, A., Compagnoni, R., Chaduteau, C., Ferraris, C.,
origin is the capability to sustain a continuous gas flow system, in terms Esteve, I., 2017. Massive production of abiotic methane during subduction evidenced
in metamorphosed ophicarbonates from the Italian Alps. Nat. Commun. 14134.
of gas emission intensity, longevity and spatial extension. In this respect, Cipolli, F., Gambardella, B., Marini, L., Ottonello, G., Zuccolini, M.V., 2004.
the CO2 hydrogenation - reservoir model, with the recognition of source Geochemistry of high-pH waters from serpentinites of the Gruppo di Voltri (Genova,
rocks and reservoir rocks, similar to sedimentary natural gas systems Italy) and reaction path modeling of CO2 sequestration in serpentinite aquifers. Appl.
Geochem. 19, 787–802.
and related seeps, can explain this aspect. The fluid inclusion and aquifer Cumming, E.A., Rietze, A., Morrissey, L.S., Cook, M.C., Rhim, J.H., Ono, S., Morrill, P.L.,
microbial models presently fail to explain how gas pressure, necessary to 2019. Potential sources of dissolved methane at the Tablelands, Gros Morne National
explain the surface emissions, can develop. Echoing Vacquand et al. Park, NL, CAN: a terrestrial site of serpentinization. Chem. Geol. 514, 42–53.
D’Alessandro, W., Yuce, G., Italiano, F., Bellomo, S., Gulbay, A.H., Yasin, D.U.,
(2018), Etiope and Whiticar (2019), and Monnin et al. (2021), metha­
Gagliano, A.L., 2018. Large compositional differences in the gases released from the
nogens collected in the wells are not necessary representative of the Kizildag ophiolitic body (Turkey): evidences of prevailingly abiogenic origin. Mar.
conditions of methane generation at depth. Microbialgenesis might be a Petrol. Geol. 89, 174–184.
surficial process, while methane may mostly originate outside the De Melo Portella, Y., Zaccarini, F., Etiope, G., 2019. First detection of methane within
chromitites of an Archean-Paleoproterozoic greenstone belt in Brazil. Minerals 9,
investigated aquifer. While CH4-bearing fluid inclusions may occur in 256. https://doi.org/10.3390/min9050256.
continental peridotites, the existence, everywhere, of large volumes of Demirel, I.H., Günay, Y., 2000. Tectonic and karstic effects on the Western Taurus
inclusions necessary to continuously sustain the various gas discharges Region, Southwestern Turkey: relations to the present temperature gradients and
total organic carbon content. Energy Sources 22, 431–441.
(bubbling seeps and high amounts of dissolved CH4 transported by Deville, E., Battani, A., Griboulard, R., Guerlais, S.H., Herbin, J.P., Houzay, J.P.,
springs) needs to be verified. For this model, there needs to be a justi­ Muller, C., Prinzhofer, A., 2003. Mud volcanism origin and processes. New insights
fication and evidence for explaining how inclusions should be spatially from Trinidad and the Barbados Prism. In: Van Rensbergen, P., Hillis, R.R.,
Maltman, A.J., Morley, C. (Eds.), Surface Sediment Mobilization, vol. 216. Special
distributed and liberated in a way to produce, for long times, the volu­ Publication of the Geological Society, London, pp. 475–490.
minous surface output of the gases observed. Additional complexity and Deville, E., Prinzhofer, A., 2016. The origin of N2-H2-CH4-rich natural gas seepages in
interpretative challenges arise when minor secondary gas sources are ophiolitic context: a major and noble gases study of fluid seepages in New Caledonia.
Chem. Geol. 440, 139–147.
involved and fluids mix during their migration to the surface. However, Economou-Eliopoulos, M., Tsoupas, G., Skounakis, V., 2019. Occurrence of graphite-like
in several study areas beyond Oman (Table S1), the origin of methane carbon in podiform chromitites of Greece and its genetic significance. Minerals 9,
can be clarified or confirmed by adding some missing analyses following 152. https://doi.org/10.3390/min9030152.
Ellison, E.T., Templeton, A.S., Zeigler, S.D., Mayhew, L.E., Kelemen, P.B., Matter, J.M.,
the approach proposed in Fig. 5 (e.g., radiocarbon in the springs in
Oman Drilling Project Science Party, 2021. Low-temperature hydrogen formation
California and Canada; fluid inclusions in the ophiolitic rocks at the during aqueous alteration of serpentinized peridotite in the Samail ophiolite.
seeps in Turkey and Japan; microbial analysis in the Ronda ophiolite J. Geophys. Res. Solid Earth 126, e2021JB021981. https://doi.org/10.1029/
springs in Spain). 2021JB021981.
Etiope, G., Baciu, C., Schoell, M., 2011b. Extreme methane deuterium, nitrogen and
helium enrichment in natural gas from the Homorod seep (Romania). Chem. Geol.
Declaration of competing interest 280, 89–96.
Etiope, G., Feyzullayev, A., Baciu, C.L., 2009. Terrestrial methane seeps and mud
volcanoes: a global perspective of gas origin. Mar. Petrol. Geol. 26, 333–344.
The authors declare that they have no known competing financial https://doi.org/10.1016/j.marpetgeo.2008.03.001.
interests or personal relationships that could have appeared to influence Etiope, G., Ifandi, E., Nazzari, M., Procesi, M., Tsikouras, B., Ventura, G., Steele, A.,
the work reported in this paper. Tardini, R., Szatmari, P., 2018. Widespread abiotic methane in chromitites. Sci. Rep.
8, 8728. https://doi.org/10.1038/s41598-018-27082-0.
Etiope, G., Ionescu, A., 2015. Low-temperature catalytic CO2 hydrogenation with
Acknowledgments geological quantities of ruthenium: a possible abiotic CH4 source in chromitite-rich
serpentinized rocks. Geofluids 15, 438–452.
Etiope, G., Samardžić, N., Grassa, F., Hrvatović, H., Miošić, N., Skopljak, F., 2017.
This work has benefitted from discussions with Edward Young, Methane and hydrogen in hyperalkaline groundwaters of the serpentinized Dinaride
Naizhong Zhang and Stefano Salvi. ophiolite belt, Bosnia and Herzegovina. Appl. Geochem. 84, 286–296. https://doi.
org/10.1016/j.apgeochem.2017.07.006.
Etiope, G., Schoell, M., 2014. Abiotic gas: atypical but not rare. Elements 10, 291–296.
Appendix A. Supplementary data
Etiope, G., Schoell, M., Hosgormez, H., 2011a. Abiotic methane flux from the Chimaera
seep and Tekirova ophiolites (Turkey): understanding gas exhalation from low
Supplementary data to this article can be found online at https://doi. temperature serpentinization and implications for Mars. Earth Planet Sci. Lett. 310,
org/10.1016/j.apgeochem.2022.105373. 96–104.
Etiope, G., Sherwood Lollar, B., 2013. Abiotic methane on Earth. Rev. Geophys. 51,
276–299. https://doi.org/10.1002/rog.20011.

11
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

Etiope, G., Tsikouras, B., Kordella, S., Ifandi, E., Christodoulou, D., Papatheodorou, G., Miura, M., Arai, S., Mizukami, T., 2011. Raman spectroscopy of hydrous inclusions in
2013a. Methane flux and origin in the Othrys ophiolite hyperalkaline springs, olivine and orthopyroxene in ophiolitic harzburgite: implications for elementary
Greece. Chem. Geol. 347, 161–174. https://doi.org/10.1016/j. processes in serpentinization. J. Mineral. Petrol. Sci. 106, 91–96. https://doi.org/
chemgeo.2013.04.003. 10.2465/jmps.101021d.
Etiope, G., Vadillo, I., Whiticar, M.J., Marques, J.M., Carreira, P.M., Tiago, I., Monnin, C., Quéméneur, M., Price, R., Jeanpert, J., Maurizot, P., Boulart, C., et al., 2021.
Benavente, J., Jiménez, P., Urresti, B., 2016. Abiotic methane seepage in the Ronda The chemistry of hyperalkaline springs in serpentinizing environments: 1. The
peridotite massif, southern Spain. Appl. Geochem. 66, 101–113. composition of free gases in New Caledonia compared to other springs worldwide.
Etiope, G., Vance, S., Christensen, L.E., Marques, J.M., Ribeiro da Costa, I., 2013b. J. Geoph. Res. Biogeosci. 126, e2021JG006243 https://doi.org/10.1029/
Methane in serpentinized ultramafic rocks in mainland Portugal. Mar. Petrol. Geol. 2021JG006243.
45, 12–16. https://doi.org/10.1016/j.marpetgeo.2013.04.009. Morrill, P.L., Brazelton, W.S., Kohl, L., Rietze, A., Miles, S.M., Kavanagh, H., Schrenk, M.
Etiope, G., Whiticar, M.J., 2019. Abiotic methane in continental ultramafic rock systems: O., Ziegler, S.E., Lang, S.Q., 2014. Investigations of potential microbial
towards a genetic model. Appl. Geochem. 102, 139–152. https://doi.org/10.1016/j. methanogenic and carbon monoxide utilization pathways in ultra-basic reducing
apgeochem.2019.01.012. springs associated with present-day continental serpentinization: the Tablelands, NL,
Fones, E.M., Templeton, A.S., Mogk, D.W., Boyd, E.S., 2022. Transformation of low CAN. Front. Microbiol. 5 https://doi.org/10.3389/fmicb.2014.00613 article 613.
molecular weight organic acids by microbial endoliths in subsurface mafic and Neal, C., Stanger, G., 1983. Hydrogen generation from mantle source rocks in Oman.
ultramafic igneous rock. Environ. Microbiol. https://doi.org/10.1111/1462- Earth Planet Sci. Lett. 66, 315–320.
2920.16041 (online version). Nivin, V.A., 2016. Free hydrogen-hydrocarbon gases from the Lovozero loparite deposit
Forzatti, P., Lietti, L., 1999. Catalyst deactivation. Catal. Today 52, 165–181. (Kola Peninsula, NW Russia). Appl. Geochem. 74, 44–55.
French, B.M., 1966. Some geological implications of equilibrium between graphite and Nothaft, D.B., Templeton, A.S., Rhim, J.H., Wang, D.T., Labidi, J., Miller, H.M., Boyd, E.
C-H-O gases at high temperatures and pressures. Rev. Geophys. 4, 223–253. S., Matter, J.M., Ono, S., Young, E.D., Kopf, S.H., Kelemen, P.B., Conrad, M.E., The
Fritz, P., Clark, I.D., Fontes, J.-C., Whiticar, M.J., Faber, E., 1992. Deuterium and 13C Oman Drilling Project Science Team, 2021. Geochemical, biological, and clumped
evidence for low temperature production of hydrogen and methane in a highly isotopologue evidence for substantial microbial methane production under carbon
alkaline groundwater environment in Oman. In: Kharaka, Y.K., Maest, A.S. (Eds.), limitation in serpentinites of the Samail Ophiolite, Oman. J. Geoph. Res. Biogeosci.
Proceed. 7th Intern. Symp. On Water–Rock Interaction: Low Temperature 126, e2020JG006025 https://doi.org/10.1029/2020JG006025.
Environments, vol. 1. Balkema, Rotterdam, pp. 793–796. Oze, C., Sharma, M., 2005. Have olivine, will gas: serpentinization and the abiogenic
Garuti, G., Zaccarini, F., 1997. In situ alteration of platinum-group minerals at low production of methane on Mars. Geophys. Res. Lett. 32, L10203 https://doi.org/
temperature: evidence from serpentinized and weathered chromitite of the vourinos 10.1029/2005GL022691.
complex, Greece. Can. Mineral. 35, 611–626. Page, N.J., Pallister, J.S., Brown, M.A., Smewing, J.D., Haffty, J., 1982. Palladium,
Giunta, T., Young, E.D., Warr, O., Kohl, I., Ash, J.L., Martini, A., Mundle, S.O.C., platinum, rhodium, iridium and ruthenium in chromite-rich rocks from the Samail
Rumble, D., Pérez-Rodríguez, I., Wasley, M., LaRowe, D.E., Gilbert, A., Sherwood ophiolite, Oman. Can. Mineral. 20, 537–548.
Lollar, B., 2019. Methane sources and sinks in continental sedimentary systems: new Pasteris, J.D., Wanamaker, B.J., 1988. Laser Raman microprobe analysis of
insights from paired clumped isotopologues 13CH3D and 12CH2D2. Geochem. experimentally re-equilibrated fluid inclusions in olivine; some implications for
Cosmochim. Acta 245, 327–351. https://doi.org/10.1016/j.gca.2018.10.030. mantle fluids. Am. Mineral. 73, 1074–1088.
Grozeva, N.G., Klein, F., Seewald, J.S., Sylva, S.P., 2020. Chemical and isotopic analyses Porter, M.L., Jiménez-Martínez, J., Martinez, R., McCulloch, Q., Carey, J.W.,
of hydrocarbon-bearing fluid inclusions in olivine-rich rocks. Philos. Trans. Royal Viswanathan, H.S., 2015. Geo-material microfluidics at reservoir conditions for
Soc. A 378. https://doi.org/10.1098/rsta.2018.0431, 2165. subsurface energy resource applications. Lab Chip 15, 4044–4053.
Hopson, C.A., Coleman, R.G., Gregory, R.T., Pallister, J.S., Bailey, E.H., 1981. Geologic Potter, J., Konnerup-Madsen, J., 2003. A review of the occurrence and origin of
section through the Samail ophiolite and associated rocks along a Muscat-Ibra abiogenic hydrocarbons in igneous rocks. In: Petford, N., McCaffrey, K.J.W. (Eds.),
transect, Southeastern Oman Mountains. J. Geophys. Res. 86, 2527–2544. Hydrocarbons in Crystalline Rocks, vol. 214. Geological Society Special Publication,
Huang, F., Barbier, S., Tao, R., Hao, J., Garcia del Real, P., Peuble, S., Merdith, A., London, pp. 151–173. The Geological Society of London.
Leichnig, V., Perrillat, J.-P., Fontaine, K., Fox, P., Daniel, I., 2021. Dataset for H2, Proskurowski, G., Lilley, M.D., Seewald, J.S., Früh-Green, G.L., Olson, E.J., Lupton, J.E.,
CH4 and organic compounds formation during experimental serpentinization. Sylva, S.P., Kelley, D.S., 2008. Abiogenic hydrocarbon production at Lost City
Geosci. Data J. 8, 90–100. hydrothermal field. Science 319, 604–607.
Klein, F., Grozeva, N.G., Seewald, J.S., 2019. Abiotic methane synthesis and Reeves, E.P., Fiebig, J., 2020. Abiotic synthesis of methane and organic compounds in
serpentinization in olivine-hosted fluid inclusions. Proc. Natl. Acad. Sci. USA 116, Earth’s lithosphere. Elements 16, 25–31.
17666–17672. Roedder, E., 1972. Composition of Fluid Inclusions. In: Data of Geochemistry. USGS
Kohl, L., Cumming, E., Cox, A., Rietze, A., Morrissey, L., Lang, S.Q., Richter, A., Numbered Series, pp. 163 No. 40-JJ.
Suzuki, S., Nealson, K.H., Morrill, P.L., 2016. Exploring the metabolic potential of Rollinson, H., Adetunji, J., 2015. Chromite in the mantle section of the Oman ophiolite:
microbial communities in ultra-basic, reducing springs at the Cedars, CA, USA: implications for the tectonic evolution of the Oman ophiolite. Acta Geol. Sin. 89
experimental evidence of microbial methanogenesis and heterotrophic acetogenesis. (Suppl. 2), 73–76.
J. Geophys. Res. Biogeosci. 121, 1203–1220. Salvi, S., Williams-Jones, A.E., 1997. Fischer–Tropsch synthesis of hydrocarbons during
Labidi, J., Young, E., Giunta, T., Kohl, I., Seewald, J., Tang, H., et al., 2020. Methane sub-solidus alteration of the Strange Lake peralkaline granite, Quebec/Labrador,
thermometry in deep-sea hydrothermal systems: evidence for re-ordering of doubly- Canada. Geochem. Cosmochim. Acta 61, 83–99.
substituted isotopologues during fluid cooling. Geochem. Cosmochim. Acta 288, Sano, Y., Urabe, A., Wakita, H., Wushiki, H., 1993. Origin of hydrogen–nitrogen gas
248–261. https://doi.org/10.1016/j.gca.2020.08.013. seeps, Oman. Appl. Geochem. 8, 1–8.
Lyon, G., Giggenbach, W.F., Lupton, J.F., 1990. Composition and origin of the hydrogen Schoell, M., 1980. The hydrogen and carbon isotopic composition of methane from
rich gas seep, Fiordland, New Zealand. EOS Trans, 1717 V51De10. natural gases of various origins. Geochem. Cosmochim. Acta 44, 649–661.
Malmqvist, L., Kristiansson, K., 1984. Experimental evidence for an ascending microflow Schrenk, M.O., Brazelton, W.J., Lang, S.Q., 2013. Serpentinization, carbon and deep life.
of geogas in the ground. Earth Planet Sci. Lett. 70, 407–416. Rev. Mineral. Geochem. 75, 575–606.
Marques, J.M., Etiope, G., Neves, M.O., Carreira, P.M., Rocha, C., Vance, S.D., Sherwood Lollar, B., Lacrampe-Couloume, G., Slater, G.F., Ward, J.A., Moser, D.P.,
Christensen, L., Miller, A.Z., Suzuki, S., 2018. Linking serpentinization, Gihring, T.M., Lin, L.-H., Onstott, T.C., 2006. Unravelling abiogenic and biogenic
hyperalkaline mineral waters and abiotic methane production in continental sources of methane in the Earth’s deep subsurface. Chem. Geol. 226, 328–339.
peridotites: an integrated hydrogeological-bio-geochemical model from the Cabeço https://doi.org/10.1016/j.chemgeo.2005.09.027.
de Vide CH4-rich aquifer (Portugal). Appl. Geochem. 96, 287–301. Sherwood Lollar, B., Lacrampe-Couloume, G., Voglesonger, K., Onstott, T.C., Pratt, L.M.,
McCollom, T.M., 2013. Laboratory simulations of abiotic hydrocarbon formation in Slater, G.F., 2008. Isotopic signatures of CH4 and higher hydrocarbon gases from
Earth’s deep subsurface. Rev. Mineral. Geochem. 75, 467–494. Precambrian Shield sites: a model for abiogenic polymerization of hydrocarbons.
McCollom, T.M., 2016. Abiotic methane formation during experimental serpentinization Geochem. Cosmochim. Acta 72, 4778–4795.
of olivine. Proc. Natl. Acad. Sci. USA 113, 13965–13970. Sleep, N.H., Meibom, A., Fridriksson, T., Coleman, R.G., Bird, D.K., 2004. H2-rich fluids
McDermott, J.M., Seewald, J.S., German, C.R., Sylva, S.P., 2015. Pathways for abiotic from serpentinization: geochemical and biotic implications. Proc. Natl. Acad. Sci.
organic synthesis at submarine hydrothermal fields. Proc. Natl. Acad. Sci. USA 112, USA 101, 12818–12823.
7668–7672. Smith, A., Oze, C., Rossman, G.R., Celestial, A.J., 2017. Raman and FTIR spectroscopy of
Melcher, F., Grum, W., Simon, G., Thalhammer, T.V., Stumpfl, F.E., 1997. Petrogenesis of methane in olivine. In: American Geophysical Union, Fall Meeting 2017 abstract
the ophiolitic giant chromite deposits of Kempirsai, Kazakhstan: a study of solid and #vols. 43C-0545.
fluid inclusions in chromite. J. Petrol. 38, 1419–1438. Stolper, D.A., Lawson, M., Formolo, M.J., Davis, C.L., Douglas, P.M.J., Eiler, J.M., 2018.
Milkov, A.V., Etiope, G., 2018. Revised genetic diagrams for natural gases based on a The utility of methane clumped isotopes to constrain the origins of methane in
global dataset of >20,000 samples. Org. Geochem. 125, 109–120. https://doi.org/ natural gas accumulations. Geol. Soc. London, Spec. Publ. 468, 23–52. https://doi.
10.1016/j.orggeochem.2018.09.002. org/10.1144/SP468.3.
Milkov, A.V., Mohinudeen, F., Etiope, G., 2020. Geochemistry of shale gases from around Stolper, D.A., Martini, A., Clog, M., Douglas, P., Shusta, S., Valentine, D., et al., 2015.
the world: composition, origins, isotope reversals and rollovers, and implications for Distinguishing and understanding thermogenic and biogenic sources of methane
the exploration of shale plays. Org. Geochem. 143, 103997 https://doi.org/ using multiply substituted isotopologues. Geochem. Cosmochim. Acta 161, 219–247.
10.1016/j.orggeochem.2020.103997. https://doi.org/10.1016/j.gca.2015.04.015.
Miller, H.M., Chaudhry, N., Conrad, M.E., Bill, M., Kopf, S.H., Templeton, A.S., 2018. Subroto, E.A., Priadi, B., Yulian, B., 2004. Study on Gas Samples Collected from Tanjung
Large carbon isotope variability during methanogenesis under alkaline conditions. Api and Tomori Area, Sulawesi: Abiogenic, Biogenic, or Thermogenic? Buletin
Geochem. Cosmochim. Acta 237, 18–31. https://doi.org/10.1016/j. Geologi, vol. 3. Institut Teknologi Bandung, pp. 117–124.
gca.2018.06.007. Suda, K., Gilbert, A., Yamada, K., Yoshida, N., Ueno, Y., 2017. Compound–and position–
specific carbon isotopic signatures of abiogenic hydrocarbons from on–land

12
G. Etiope and C. Oze Applied Geochemistry 143 (2022) 105373

serpentinite–hosted Hakuba Happo hot spring in Japan. Geochem. Cosmochim. Acta Wang, D.T., Welander, P.V., Ono, S., 2016. Fractionation of the methane isotopologues
13
206, 201–215. CH4, 12CH3D, and 13CH3D during aerobic oxidation of methane by Methylococcus
Suda, K., Aze, T., Miyairi, Y., Yokoyama, Y., Matsui, Y., Ueda, H., et al., 2022. The origin capsulatus (Bath). Geochem. Cosmochim. Acta 192, 186–202. https://doi.org/
of methane in serpentinite-hosted hyperalkaline hot spring at Hakuba Happo, Japan: 10.1016/j.gca.2016.07.031.
radiocarbon, methane isotopologue and noble gas isotope approaches. Earth Planet Warr, O., Young, E.D., Giunta, T., Kohl, I.E., Ash, J.L., Lollar, B.S., 2021. High-resolution,
Sci. Lett. 585, 117510. long-term isotopic and isotopologue variation identifies the sources and sinks of
Szponar, N., Brazelton, W.J., Schrenk, M.O., Bower, D.M., Steele, A., Morrill, P.L., 2013. methane in a deep subsurface carbon cycle. Geochem. Cosmochim. Acta 294,
Geochemistry of a continental site of serpentinization in the Tablelands ophiolite, 315–334.
Gros Morne National Park: a Mars analogue. Icarus 224, 286–296. Whiticar, M.J., Faber, E., Schoell, M., 1986. Biogenic methane formation in marine and
Suzuki, S., Kuenen, J.G., Schipper, K., Van Der Velde, S., Ishii, S.I., Wu, A., et al., 2014. freshwater environments: CO2 reduction versus acetate fermentation - isotope
Physiological and genomic features of highly alkaliphilic hydrogen-utilizing evidence. Geochem. Cosmochim. Acta 50, 693–709.
Betaproteobacteria from a continental serpentinizing site. Nat. Commun. 5, 1–12. Wycherley, H., Fleet, A., Shaw, H., 1999. Some observations on the origins of large
Taenzer, L., Labidi, J., Masterson, A.L., Feng, X., Rumble III, D., Young, E.D., Leavitt, W. volumes of carbon dioxide accumulations in sedimentary basins. Mar. Petrol. Geol.
D., 2020. Low Δ12CH2D2 values in microbialgenic methane result from 16, 489–494.
combinatorial isotope effects. Geochem. Cosmochim. Acta 285, 225–236. Woycheese, K.M., Meyer-Dombard, D.R., Cardace, D., Arcilla, C.A., Ono, S., 2017.
Tiago, I., Veríssimo, A., 2013. Microbial and functional diversity of a subterrestrial high Metagenomic and clumped isotopologue evidence for microbial methanogenesis in
pH groundwater associated to serpentinization. Environ. Microbiol. 15, 1687–1706. the Zambales ophiolite. In: AGU Fall Meeting Abstracts 2017, pp. B31E–B2038.
Twing, K.I., Brazelton, W.J., Kubo, M.D.Y., Hyer, A.J., Cardace, D., Hoehler, T.M., et al., Xia, X., Gao, Y., 2019. Kinetic clumped isotope fractionation during the thermal
2017. Serpentinization-influenced groundwater harbors extremely low diversity generation and hydrogen exchange of methane. Geochem. Cosmochim. Acta 248,
microbial communities adapted to high pH. Front. Microbiol. 8, 308. https://doi. 252–273.
org/10.3389/fmicb.2017.00308. Xia, X., Gao, Y., 2021. Validity of geochemical signatures of abiotic hydrocarbon gases on
Ueda, H., Matsui, Y., Sawaki, Y., 2021. Abiotic methane generation via CO2 Earth. J. Geol. Soc. 23 https://doi.org/10.1144/jgs2021-077.
hydrogenation with natural chromitite under hydrothermal conditions. G-cubed 22, Young, E.D., 2020. A two-dimensional perspective on CH4 isotope clumping:
e2020GC009533. https://doi.org/10.1029/2020GC009533. distinguishing process from source. In: Orcutt, B.N., Daniel, I., Dasgupta, R. (Eds.),
Vacquand, C., Deville, E., Beaumont, V., Guyot, F., Sissmann, O., Pillot, D., Arcilla, C., Deep Carbon: Past to Present. Cambridge University Press, pp. 388–414. https://doi.
Prinzhofer, A., 2018. Reduced gas seepages in ophiolitic complexes: evidences for org/10.1017/9781108677950.
multiple origins of the H2-CH4-N2 gas mixtures. Geochem. Cosmochim. Acta 233, Young, E.D., Kohl, I., Lollar, B.S., Etiope, G., Rumble III, D., Li, S., et al., 2017. The
437–461. relative abundances of resolved 12CH2D2 and 13CH3D and mechanisms controlling
Van den Kerkhof, A.M., Hein, U.F., 2001. Fluid inclusion petrography. Lithos 55, 27–47. isotopic bond ordering in abiotic and biotic methane gases. Geochem. Cosmochim.
Wang, D.T., Gruen, D.S., Sherwood Lollar, B., Hinrichs, K.-U., Stewart, L.C., Holden, J.F., Acta 203, 235–264. https://doi.org/10.1016/j.gca.2016.12.041.
Hristov, A.N., Pohlman, J.W., Morrill, P.L., Könneke, M., Delwiche, K.B., Reeves, E. Zhang, N., Sekine, Y., Yamada, K., Nakagawa, M., Lin, M., Yoshida, N., 2021. Clumped
P., Sutcliffe, C.N., Ritter, D.J., Seewald, J.S., McIntosh, J.C., Hemond, H.F., Kubo, M. isotope signatures of abiotic methane formed via nickel-catalyzed Fischer-Tropsch
D., Cardace, D., Hoehler, T.M., Ono, S., 2015. Nonequilibrium clumped isotope synthesis and their implications. In: AGU 2021 Fall Meeting Abstract. https://agu
signals in microbial methane. Science 348, 428–431. https://doi.org/10.1126/ 2021fallmeeting-agu.ipostersessions.com/Default.aspx?s=96-E1-B4-70-42-37-97-
science.aaa4326. AC-6B-7A-8D-18-C1-44-A9-52.
Wang, D.T., Reeves, E.P., McDermott, J.M., Seewald, J.S., Ono, S., 2018. Clumped Zwicker, J., Birgel, D., Bach, W., Richoz, S., Smrzka, D., Grasemann, B., Gier, S.,
isotopologue constraints on the origin of methane at seafloor hot springs. Geochem. Schleper, C., Rittmann, S.K.-M.R., Kosun, E., Peckmann, J., 2018. Evidence for
Cosmochim. Acta 223, 141–158. https://doi.org/10.1016/j.gca.2017.11.030. archaeal methanogenesis within veins at the onshore serpentinite-hosted Chimaera
Wang, W., Wang, S., Ma, X., Gong, J., 2011. Recent advances in catalytic hydrogenation seeps, Turkey. Chem. Geol. 483, 567–580.
of carbon dioxide. Chem. Soc. Rev. 40, 3703–3727.

13

You might also like