Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Earth-Science Reviews 230 (2022) 104063

Contents lists available at ScienceDirect

Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev

Molecular hydrogen in surface and subsurface natural gases: Abundance,


origins and ideas for deliberate exploration
Alexei V. Milkov
Colorado School of Mines, Golden, CO, USA

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

Keywords: Geologic molecular hydrogen (H2) occurs in the subsurface and vents and seeps at the surface. However, this
Hydrogen valuable natural resource is under-utilized in the economy because the distribution, abundance and origins of H2
Serpentinites are poorly understood. I studied a global dataset of 6246 natural gases with reported H2 concentrations from 16
Natural gas
different geological habitats. The average H2 concentration in all gas samples is 3.5%, but the median concen­
Isotopes
tration is only 0.01%. Gases sampled in Mid-Ocean Ridges and in serpentinites have the highest average con­
centrations of H2 (~24% and ~21%, respectively). More than 30 different processes may produce H2 observed in
natural gases. Hydrogen isotopic composition (expressed as δ2H-H2 values) may indicate crustal (<-650‰) or
mantle and primordial (from -650‰ to -100‰) sources of H2, or may result from temperature-dependent
equilibration of H2 with water. Much of crustal H2 may be sourced by the reactions of serpentinization, while
the quantitative significance of other H2-generating processes such as radiolytic decomposition of water and
hydrocarbons, fracture-induced reduction of water, petroleum cracking and coal metamorphism remain specu­
lative. Primordial H2 perhaps vents in some volcanic settings. Provided better understanding of H2 abundance
and origins in different geological settings should enable the purposeful exploration for geologic H2 and the
assessment of its economic resources.

1. Introduction surface (Zgonnik, 2020; Etiope et al., 2011; Combaudon et al., 2022).
This geologic (natural, native, gold) H2 is already utilized as a source of
Hydrogen, the most abundant element in the Universe, was recog­ energy (electricity) in a village in Mali (Hydroma, 2022), and should be
nized in 1776 by Henry Cavendish, and its name was coined in 1783 by used more widely around the globe if available, producible and
Antoine Lavoisier (Brophy and Schimmelmann, 2017). Molecular economically viable (Smith, 2002; Larin et al., 2015; Truche and
hydrogen (H2) is important in modern economy and industry. Currently, Bazarkina, 2019; Gordienko, 2019; Gaucher, 2020; Boreham et al.,
it is used mostly in ammonia production and petroleum recovery and 2021a; Lefeuvre et al., 2021; Yedinak, 2022).
refining (IEA, 2021). Hydrogen gas is an excellent clean fuel because it Hydrogen and other reduced gases support unique ecosystems with
contains more energy per unit weight than most other fuels, is carbon- chemolithoautotrophic metabolic pathways on the modern deep sea­
free, and its combustion in fuel cells emits only water (World Energy floor (Takai et al., 2004; Ohara et al., 2012), may provide energy for
Council, 2019; Truche and Bazarkina, 2019; Farias et al., 2022). long-term survival of microbial cells (Morita, 2000) and may have
Therefore, H2 can play an important role in the energy transition away contributed to the origin of life on Earth and perhaps other planets
from fossil fuels as either an energy vector or a zero-carbon source of (Nisbet and Sleep, 2001; Oze and Sharma, 2005, 2007; McCollom and
energy (Hanley et al., 2018; Murray et al., 2020). The vast majority of Seewald, 2013; Holm et al., 2015; Glein and Zolotov, 2020; McCollom
the H2 currently utilized in the economy is produced using methods et al., 2022). Molecular H2 plays a significant role in the formation of
collectively known as “thermal processes”, and steam reforming of natural hydrocarbon gases via the processes of microbial hydrogeno­
methane and other light hydrocarbons is the most common industrial trophy (Wagner et al., 2018; Katz, 2011; Milkov, 2011, 2018) and
process of H2 generation (Nuttall and Bakenne, 2020; Timmerberg et al., Fischer-Tropsch synthesis (Etiope and Whiticar, 2019), and may play a
2020). role in the formation of liquid petroleum fluids (Pratt, 1934; Szatmari,
Molecular H2 occurs in the subsurface and seeps and vents at the 1989; Levshounova, 1991; Kenney et al., 2002; Wang et al., 2019).

E-mail address: amilkov@mines.edu.

https://doi.org/10.1016/j.earscirev.2022.104063
Received 29 January 2022; Received in revised form 16 May 2022; Accepted 22 May 2022
Available online 28 May 2022
0012-8252/© 2022 Elsevier B.V. All rights reserved.
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Emissions of combustible and explosive H2-containing gases during the 2.2.1. Natural abiotic processes
mining of ore deposits represent a known hazard (Nivin, 2019) and need Natural abiotic processes of H2 generation in Earth include water-
to be predicted and mitigated. rock interactions involving iron, degassing from the mantle, radiolysis
In spite of the clear importance of geologic H2 as a clean fuel, a and other abiotic processes in the geosphere:
source of energy for life and a geohazard, the current understanding of 1. Serpentinization (metamorphic hydration and oxidation) of ul­
geologic H2 occurrences is rather limited, and the in-place, recoverable tramafic rocks such as peridotites, dunites and kimberlites composed
and economic resources of geologic H2 have not been adequately mostly of Mg and Fe2+ silicate minerals olivine and pyroxene (Fridman,
quantified. The main reason for that is the lack of fundamental under­ 1970; Moody, 1976; Coveney Jr. et al., 1987; Sleep et al., 2004; Oze and
standing of the prevailing reaction mechanisms and rates associated Sharma, 2005, 2007; Russell et al., 2010; Klein et al., 2013; Mayhew
with geochemical processes of H2 formation, and how H2 migrates and et al., 2013; Holm et al., 2015; Miller et al., 2017; Preiner et al., 2018;
fractionates in the Earth. More than 30 different mechanisms and pro­ Huang et al., 2019; Truche et al., 2020; Worman et al., 2016, 2020;
cesses of H2 production in Earth have been proposed in the literature. Zgonnik, 2020; Albers et al., 2021). Serpentinites are water-rich rocks
However, the relative importance of these processes and their contri­ that contain mostly serpentine-group minerals (chrysotile, lizardite and
butions to global H2 budget remain unknown. This is partly due to the antigorite) and varying amounts of brucite, magnetite, and/or FeNi al­
fact that geochemical tools aimed at distinguishing these processes are loys (Sleep et al., 2004). They form from relatively low-temperature
poorly developed. Further, various different processes that consume H2 (<300◦ C) hydration of mantle-derived ultramafic rocks, during which
in the subsurface and lithologies that can form effective seals for H2- Mg-rich olivine and orthopyroxene are altered to serpentine minerals
bearing reservoirs are not well understood. (Guillot and Hattori, 2013; Evans et al., 2013).
Here, I use a large global dataset of natural gases to document the Modern serpentinites in oceans are present along the axis of slow and
abundance of molecular H2 in various geological settings. Then, I very slow spreading ocean ridges where the mantle is too cool to form
investigate how to distinguish numerous proposed processes of H2 significant basalt (Charlou et al., 2002; Dick et al., 2003; Worman et al.,
generation in the subsurface and which of them may be quantitatively 2020), magma-poor transform margins (such as the Iberian Margin,
significant on Earth by using H2 isotopic composition and other asso­ Skelton et al., 2005; Albers et al., 2021) and in the subduction zones
ciated geological and geochemical characteristics (including noble (some forearc environments, for example, in the serpentinite diapirs in
gases) of H2-bearing gases. A better understanding of processes that the Mariana forearc, Wang et al., 2009; Fryer, 2012). Ancient serpen­
generate H2 will help constrain the rates of geologic hydrogen produc­ tinites on continents are exposed as ophiolites in suture zones associated
tion in Earth, which may be significant (up to ~30% of global industrial with closure of paleo-oceans. The largest (by area) examples include the
hydrogen production, Sherwood Lollar et al., 2014) but are highly un­ Semail ophiolites (~10,000 km2) in the Sultanate of Oman and the
certain (Zgonnik, 2020). It will also lay out the foundation for the United Arab Emirates (Barnes et al., 1978; Neal and Stranger, 1983; Fritz
assessment of potentially recoverable resources of geologic H2 around et al., 1992; Kelemen and Matter, 2008) and the Massif du Sud (~6000
the world and help formulate exploration strategies for this valuable gas. km2) in New Caledonia (Barnes et al., 1978; Deville and Prinzhofer,
2016; Monnin et al., 2021). Ophiolites with smaller areas are present in
2. Current knowledge of hydrogen occurrences, origins and New Zealand (Wood, 1972), Philippines (Thayer, 1966; Abrajano et al.,
sinks 1988; Sturchio et al., 1989), Japan (Suda et al., 2014, 2017), China (Ye
et al., 2007), Turkey (Etiope et al., 2011; Yuce et al., 2014; D’Alessandro
2.1. Occurrences of molecular H2 on Earth et al., 2018), Bosnia and Herzegovina (Barnes et al., 1978; Etiope et al.,
2017), Greece (Daskalopoulou et al., 2018; Li Vigni et al., 2021), Italy
The H2 molecule has a small size, is extremely mobile and can be (Boschetti et al., 2013), Spain (Etiope et al., 2016), Portugal (Etiope
rapidly consumed by redox reactions mediated by microbes and/or et al., 2013; Marques et al., 2018), Norway (Daae et al., 2013), Mali
mineral catalysts (Brophy and Schimmelmann, 2017; Truche et al., (Caby, 2014), Costa-Rica (Crespo-Medina et al., 2017), USA (Morrill
2018). As a result, H2 does not accumulate easily in the crust, which led et al., 2013; Lazar and Cooperdock, 2021), Canada (Szponar et al., 2013)
to a misconception that it does not occur freely in the subsurface (Smith and in other countries, on all continents, including Antarctica. Guillot
et al., 2005). However, several recent reviews (Gregory et al., 2019; and Hattori (2013) estimated that >3% of the Earth’s surface is made up
Truche et al., 2020; Zgonnik, 2020; Rusakov, 2020) highlighted that of serpentinites and serpentinized peridotite.
concentrations of molecular H2 can be very high (up to 99 vol.%) in Serpentinization of peridotite is accompanied by the release of H2
ophiolites and associated seeps (e.g., Neal and Stranger, 1983; Vacquand and other reduced gases. Olivine is the main mineral of the upper
et al., 2018) and in kimberlite pipes (e.g., Fridman, 1970, also note that mantle. The general reaction that describes serpentinization of olivine is
H2 is a major gas released from diamonds, Melton and Giardini, 1974). (McCollom et al., 2022):
Other geologic settings with significant H2 occurrences include graphite Olivine + H2O → Serpentine ± Brucite ± Magnetite + H2
deposits, volcanic systems, geothermal systems, crystalline basement, Generation of H2 during serpentinization reactions has been
potash and evaporite deposits, cataclasites, and anoxic sediments demonstrated in laboratory experiments (Miller et al., 2017; Huang
(Gregory et al., 2019; Zgonnik, 2020) as well as some conventional oil et al., 2019; Barbier et al., 2020; Song et al., 2021). Serpentinization is
and gas accumulations (Bohdanowicz, 1934; Headlee, 1962; Molcha­ thought to be the major process responsible for some of the highest
nov, 1981; Angino et al., 1984; Coveney Jr. et al., 1987; Perevozchikov, concentrations of H2 in natural gas seeps (Abrajano et al., 1988; Etiope
2011; Zinger, 1962; Kraychuk, 1976; Prinzhofer et al., 2018). et al., 2013) and in subsurface gas reservoirs (Coveney Jr. et al., 1987;
Briere et al., 2017). Serpentinization of iron-rich olivine on Mars, icy
moons, and other planetary bodies may be a significant source of
2.2. Origins of molecular H2 in natural gases hydrogen, which has a potential to support extra-terrestrial biological
communities (McCollom et al., 2022).
There are numerous proposed origins and generation mechanisms of 2. Release of H2 during volcanic activity and from hydrothermal
H2 observed in the subsurface and in seeps/vents at the surface and the vents, when equilibrium reactions in the C-H-O-S system in magmas
ocean floor, as previously reviewed by Lin et al. (2005a, 2005b), Truche favor H2, for example (Apps and van de Kamp, 1993; Holland, 2002):
et al. (2020), Klein et al. (2020), Zgonnik (2020) and Boreham et al. CO + H2O ↔ CO2 +H2
(2021a). I identified 32 processes previously described in the literature H2S + 2H2O ↔ SO2 + H2
and list them below grouped as natural abiotic, natural biotic, and SO2 + 2H2O ↔ H2SO4 + H2
anthropogenic processes. 3. Cooling and decompression in a hydrothermal CO2-CH4-C fluid

2
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

system (Tsunogae and Dubessy, 2009). hydrocarbons (Boreham and Davies, 2020).
4. Late stage crystallization of basic magma where dissolved water 16. Fracture-induced reduction of water: cataclastics (cohesive
oxidizes ferrous iron Fe2+ (in Fe2+-bearing minerals) to ferric iron Fe3+ granular faulted rock) of silicates in the presence of water and in igneous
(Stevens and McKinley, 1995; Schrenk et al., 2013; Sherwood Lollar rocks and associated with upper crustal processes (e.g., earthquakes)
et al., 2014; Etiope and Whiticar, 2019; Nivin, 2019). An example re­ (Kita et al., 1982; Sugisaki et al., 1983; Apps and van de Kamp, 1993; Ito
action would be decomposition of fayalite without serpentinization et al., 1999; Freund et al., 2002; Saruwatari et al., 2004; Kameda et al.,
(Arrouvel and Prinzhofer, 2021): 2004; Hirose et al., 2011, 2012; Klein et al., 2020). The mechanism is
3Fe2SiO4 (fayalite) + 2H2O(l) → 2Fe3O4 (magnetite) + 3SiO2(­ likely mechanoradical where 2(≡Si⋅) + 2H2O → 2(≡SiOH) + H2, with
quartz) + 2H2(g) the potential for subsequent redox conversion of hydroxyl (2(≡Si–OH)
5. Degassing of primordial H2 that has been present in the Earth since or 2(–OH)) to peroxy links (≡SiOOSi≡ (or -OO-)) + H2.
its formation (Chamberlin, 1909; Larin, 1993; Gilat and Vol, 2005, 17. Mechanochemical reaction of phosphate to produce phosphine
2012; Larin et al., 2015; Yang et al., 2016). (PH3) and its subsequent hydrolysis to H2 (Hoeman, 1944; Glindermann
6. Metasomatism of upper mantle and crustal rocks with the infil­ et al., 2005).
tration of metal hydrides from the deeper mantle (Walshe, 2006). 18. Natural release of H2 from fluid inclusions. This process appears
7. Hydration of biotite in felsic rocks such as granite (Murray et al., to be widespread in Precambrian basement rocks, particularly granites.
2020). The eroded basement remnants in sedimentary basins may be a signif­
8. Hydrothermal hydration of siderite in sedimentary basins (e.g., in icant H2 resource (Parnell and Blamey, 2017a).
the Solimões basin in Brazil, Milesi et al., 2016) and metasedimentary 19. Decomposition of methane (CH4) to form graphite and H2 at high
rocks according the following reaction (Seewald, 2001; Milesi et al., (>600◦ C) temperatures during metamorphism (Galimov et al., 1973;
2015): Hahn-Weinheimer and Hirner, 1981; Apps and van de Kamp, 1993;
3FeCO3 (siderite) + H2O (l) ↔ Fe3O4 (magnetite) + 3CO2 (g) + H2 Holland, 2002; Smith et al., 2005; Dzaugis et al., 2016; Bouquet et al.,
(g) 2017; Parnell and Blamey, 2017a, 2017b; Truche et al., 2018; Boreham
9. Oxidation of Fe2+ from magnetite to Fe3+ to crystallize hematite and Davies, 2020; Zgonnik, 2020):
(Arrouvel and Prinzhofer, 2021): CH4 ↔ C(graphite) + H2
2Fe3O4 (magnetite) + H2O(l) → 3Fe2O3 (hematite) + H2 (g) That CH4 may be biotic or abiotic, but the formation of H2 is an
Geymond et al. (2022) suggested that weathering of banded iron abiotic process. Hydrogen can also be generated from carbonaceous
formations can result in significant H2 generation in iron mines in material during graphitization (Suzuki et al., 2017).
Australia, Brazil and South Africa. 20. Oxidative coupling of methane catalyzed by alkali metal and
10. Reactions linked to pyritization. Under acidic conditions (such as lanthanide oxides. This process has been demonstrated in the laboratory
in some hydrothermal sites, fumaroles from volcanic activities), the at high (>650◦ C) temperatures, and similar reactions may potentially
reactions are simulated with H2S (g) (Arrouvel and Prinzhofer, 2021): occur in the subsurface with the involvement of O2 from water radiolysis
Fe2O3 (hematite) + 4H2S (g) → 2FeS2 (pyrite) + 3H2O (l) + H2 (g) (Lunsford, 1990; Lehmann and Baerns, 1992).
Fe3O4 (magnetite) + 6H2S (g) → 3FeS2 (pyrite) + 4H2O(l) + 2H2(g) 21. Mixing of waters with differing ionisation potentials, i.e. fresh
Reactions may also be with siderite and with pyrrhotite, which may and saline groundwaters (Goebel et al., 1984).
also be mediated by microorganisms (Drobner et al., 1990; Hofstra and 22. Release from ammonium (NH+ 4 )-bearing minerals oxidized by
Cline, 2000; Thiel et al., 2019). reacting with sulfate at high temperatures (Dubessy et al., 1989; Li et al.,
11. Water hydrolysis accompanying Fe and Fe2+ oxidation (anaer­ 2009), which may happen in anhydrite-rich marine salt deposits
obic corrosion) most likely catalysed by metal hydroxides at low tem­ (Dubessy and Ramboz, 1986):
2-
perature, crustal weathering and by Ca produced from the radioactive 8NH+ 4 + 3SO4 ↔ 4N2 + 3H2S + 12H2O + 2H
+

decay of 40K (Neal and Stranger, 1983; Guélard et al., 2017; Worman 23. Dehydrogenation of clay minerals that are abundant in many
et al., 2020). shale formations (Truche et al., 2018).
12. Radiolysis of water by alpha, beta and gamma radiation released 24. Liberation of H2 from H2O at high metamorphic temperatures
in the radioactive decay of U-, Th- and K-bearing minerals (Boyd et al., (Suzuki et al., 2017).
1973; Vovk, 1987; Dubessy et al., 1988, 1989; Spinks and Woods, 1990;
Smith et al., 2005; Lin et al., 2005a, 2005b; Blair et al., 2007; Dzaugis 2.2.2. Natural biotic processes
et al., 2016; Bouquet et al., 2017; Parnell and Blamey, 2017b; Truche Natural biotic processes (both biologically-mediated processes and
et al., 2018; Wang et al., 2019; Zgonnik, 2020; Sauvage et al., 2021). processes that may be abiotic but which act upon biogenic substrates) of
When ionizing radiation passes through water, it leads to the formation H2 generation include:
of ionic and excited states, which further decompose or recombine to 25. Formation during biological activity (Ginsburg-Karagitscheva,
produce radical and molecular species according to the following 1933; ZoBell, 1947; Bokova, 1953; Ekzercev, 1960; Nandi and Sengupta,
equations (Pastina and LaVerne, 2001; Wang et al., 2019): 1998; Gregory et al., 2019) via fermentation of organic matter (Gest,
H2O (ionizing radiation) → H⋅ + OH⋅ + H+ + OH− + e− 1954; Krichevsky et al., 1961; Adams and Stiefel, 1988; Boone et al.,
e− + → e−aq (hydration) 1989; Adam and Perner, 2018; Hallenbeck and Benemann, 2002) by
e−aq + H+ → H⋅ certain anaerobic bacteria and cyanobacteria (Luo et al., 1991; Conrad,
e−aq + e−aq +2H2O → H2 + 2OH− 1996; Morita, 2000; Ménez, 2020), by heterotrophic hyperthermophiles
e−aq + H⋅ + H2O → H2 + OH− (Mermelstein and Zeikus, 1998), by nitrogen fixing bacteria (Morita,
H⋅ + H2O → H2 + OH⋅ 2000; Conrad, 1996), and through anaerobic carbon monoxide oxida­
H⋅ + H⋅ → H2 tion and phosphate oxidation (Schwartz et al., 2013; Sipma et al., 2006).
13. Radiolytic dehydrogenation of organic matter and kerogen, for During diagenesis, the polymerization of hydrocarbon C-H bonds results
example, in shales, which generates gases dominated by H2 and CO2 in the formation of kerogen and release of H2.
(Lewan et al., 1991; Wang et al., 2022). The generation of H2 appears to 26. Thermogenic liberation of H2 from cracking of C-H bonds by
be due to C-H scission and double-bond formation (Yamazaki and Shida, aromatization, polycondensation or metathesis of hydrocarbons (Li
1960; Frolov et al., 1998). et al., 2015) during the generation of oil and gas (Bogomolov, 1976;
14. Radiolytic dehydrogenation of crude oil molecules (Frolov et al., Gregory et al., 2019, and references therein), in oil accumulations
1998; Larter et al., 2019; Silva et al., 2019). (Zinger, 1962), during petroleum cracking (Briere et al., 2017) and
15. Radiation-induced polymerization of methane into H2 and C2+ during coal metamorphism (Zgonnik, 2020). During metagenesis, the

3
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

progressive aromatisation of the residual kerogen leads to production unsaturated) rocks, at relatively low temperatures, in the presence of
predominantly of CH4 but also some H2 (Tissot and Welte, 1984; Hunt, minerals that can act as FTT catalysts (e.g., chromium and/or
1996; Horsfield et al., 2022). Suzuki et al. (2017) found significant ruthenium-based minerals in ophiolites).
concentrations of presumably organic H2 in gases released from pul­ There may be many inorganic sinks of H2 in the subsurface (reactions
verized thermally mature shales and matapelites. Liberation of molec­ with pyrite or other sulphides, oxidation to water and others). For
ular H2 from shales and coals during laboratory pyrolysis experiments example, Suzuki et al. (2017) suggested that the dehydration of hydrous
was demonstrated in numerous studies (Jüntgen and van Heek, 1979; minerals could be responsible for consumption of H2 during
Campbell et al., 1980; Li et al., 2015, 2017; Horsfield et al., 2022). metamorphism.
However, the specific mechanisms of H2 generation (cracking of hetero- Some geologic H2 escapes into the atmosphere and the ocean, for
bonds, demethylation, aromatization, condensation) and the role of H2 example, in areas where H2-rich gases seep at the surface (Etiope et al.,
in the transformation of organic matter into hydrocarbons are still 2011) and vent at the seafloor (Charlou et al., 2002). However, the
poorly constrained (Behar et al., 1997; Li et al., 2015, 2017). models and budgets of H2 in the atmosphere currently do not account for
27. Thermal decomposition of alkanes and carboxylic acids at tem­ the geologic H2 (Price et al., 2007; Ehhalt and Rohrer, 2009; Pieterse
peratures >200◦ C and CH4 at temperatures >500◦ C, which results in the et al., 2011). Although H2 is commonly considered as
release of H2 (Tissot and Welte, 1984; Seewald, 2001). environment-friendly gas (World Energy Council, 2019), its natural or
28. H2 release during sulfur incorporation into hydrocarbon com­ industrial combustion adds water vapor to the atmosphere, which is the
pounds during thermochemical sulfate reduction (Ostertag-Henning and most abundant and powerful greenhouse gas (Sherwood et al., 2018). In
Scheeder, 2009). addition, leakage of H2 into the atmosphere reduces global levels of the
hydroxyl radical OH and therefore increases the lifetime and forcing of
2.2.3. Anthropogenic processes the second most powerful greenhouse gas CH4 (Warwick et al., 2004). A
Anthropogenic processes in which H2 results from artificial processes recent study estimated that, over a 100-year time period, one metric ton
during drilling and metal corrosion include: of H2 in the atmosphere will warm the Earth some 11 (± 5) times more
29. Drill bit metamorphism, bit burning and cracking of organic than one metric ton of CO2 (Warwick et al., 2022). An increase in the
matter by excessively hot drill bit (Halas et al., 2021). atmospheric H2 also increases the concentration tropospheric ozone
30. The oxidation of steel drill pipes and well casings in the presence (O3), important to both air quality and climate (Warwick et al., 2022).
of natural gas with a high CO2 content (“CO2 corrosion”) (Meincke,
1967; Mørkved et al., 2009). 3. Dataset
31. Corrosion of steel well casing from exposure to H2S (Lyon and
Hulston, 1984). Compilation of a large global dataset of ~34,500 gas samples from
32. Corrosion reactions between acids either naturally present in 100 countries and territories across seven continents (~370 basins) was
ground waters at low pH conditions (Zinger, 1962) or introduced during done using data from ~1200 publications and public databases. Gas
wellbore cleaning operations (Feldbusch et al., 2018) and iron in well samples were taken at depths ranging from the surface to 8600 m below
casings, steel drilling shavings and lithic fragments (Bjornstad et al., surface/seafloor. Gas samples came from 16 different geological habitats
1994), for example: listed below (note that the references cited below are not the sources of
2H3O+ + Fe → 2H2O + H2 + Fe2+ all of the data from the given habitat but rather provide further details
When nuclear waste is disposed, corrosion of steel canisters can about the gas type, collection method or source of the samples typical for
produce copious amounts of hydrogen (Truche et al., 2020). Other ref­ that habitat).
erences describing these anthropogenic processes include Goebel et al.
(1984), Merlivat et al. (1987), Zimmer (1993), Zimmer and Erzingeer A. Conventional oil and gas reservoirs. These reservoirs do not
(1995), Messer et al. (2008), Wersin et al. (2008), Erzinger et al. (2006), require stimulation such as fracking or heating to recover oil and
Wenger et al. (2009), Thiessen et al. (2009). gas. The gases were sampled either from wellheads or separators
at the surface (Liu et al., 2013) or directly from producing for­
2.3. Sinks of H2 on Earth mations (mostly with formation testers such as Modular Dynamic
Formation Tester, Milkov et al., 2007; Petersen et al., 2019).
Based on the available evidences, it appears that virtually all of the B. Coals. These gases came from coal mines and coal seams (known
H2 generated in the subsurface is converted to methane (and other hy­ as coal bed methane (CBM) or coal seam gas (CSG)) (Milkov,
drocarbons) via biotic and abiotic processes (Conrad, 1995). Hydro­ 2021). They were sampled from producing wellheads (Bao et al.,
genotrophic microbes (Wagner et al., 2018; Adam and Perner, 2018) 2014) or from degassed coals placed in desorption canisters
utilize H2 and CO2 to produce primary microbial gas (mostly methane) (Bannerjee et al., 2021).
in marine sediments (e.g., Katz, 2011) and secondary microbial gas in C. Shales. These gases came from geological formations with true
shallow cool petroleum reservoirs (Milkov, 2010, 2011, 2018). shale lithology such as the Marcellus Formation in the USA and
Abiotic methane (Potter and Konnerup-Madsen, 2003; Etiope and the Wufeng-Longmaxi Formation in China and from unconven­
Sherwood Lollar, 2013; Etiope and Whiticar, 2019) forms from H2 via tional “tight” reservoirs composed of very fine-grained sand­
Fischer-Tropsch Type reactions (FTT, Fischer and Tropsch, 1926). The stone/siltstone (e.g., the Bakken Formation, USA) and some
FTT chemical reactions have been widely studied and applied in in­ mixed clastic/carbonate reservoirs (e.g., the Niobrara Formation,
dustrial catalysis for synthetic fuel and chemicals production and CO2 USA) where gases were generated within these formations or in
consumption (Anderson, 1984; Schulz, 1999; Wang et al., 2011). The immediately surrounding rocks (Milkov et al., 2020a). The gases
general form of the FTT reaction is: were sampled during production at wellheads (Feng et al., 2017)
(2n + 1)H2 + nCO = CnH2n+2 + nH2O or after degassing from rocks placed in desorption jars and can­
The highly exothermic hydrogenation of CO is a basic FTT reaction isters (Chatellier et al., 2013; Chen et al., 2020).
(Rӧper, 1983; Szatmari, 1989). The Sabatier reaction of slightly endo­ D. Hydrocarbon seeps and mud volcanoes. These gases were
thermic CO2 hydrogenation is another form of FTT (Sabatier and sampled at the surface/mudline from actively or passively
Senderens, 1902). In this process, CO is formed first, and is then hy­ degassing petroleum seeps and mud volcanoes in sedimentary
drogenated to hydrocarbons (Weatherbee and Bartholomew, 1984). basins (Etiope et al., 2009).
Etiope and Whiticar (2019) suggested that most abiotic methane forms E. Marine sediments. These gases were sampled in marine sediments
from CO2 hydrogenation (Sabatier reaction) in water-free (or with no obvious indication of petroleum or volcanic/magmatic

4
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

seeps or discharge from hydrothermal systems, and recovered by O. Springs. These gases were sampled at the surface/mudline from
piston coring (Bernard et al., 1977, 2013) or during ocean drilling some degassing features that do not have obvious known asso­
(Choi et al., 2016). Such gases are usually water-dissolved in the ciation with either hydrocarbon seeps (and underlying petroleum
subsurface, and were sampled after being released into gas voids systems) or with degassing volcanoes associated with magmatic
(Milkov et al., 2004) or into headspace (Luong et al., 2021). processes (and having obvious thermal anomalies). Examples
F. Gas hydrates. These gases were sampled during the decomposi­ include springs along the Dead Sea Fault system (Gasperini et al.,
tion of gas hydrates recovered from marine (Milkov, 2005) and 2020). Many of the springs from the database of Mariner et al.
continental (Liu et al., 2020; Hachikubo et al., 2020) sediments. (2006) were included in this category simply because there was
G. Continental sediments (including lakes and swamps). These gases no clear geological information to classify them as either volca­
were sampled in continental settings (Hendrix, 2009) in sedi­ nic/magmatic seeps or hot springs in sedimentary hydrothermal
ments and rocks with no obvious indication of petroleum or system.
volcanic/magmatic seeps or discharge from hydrothermal sys­ P. Serpentinites. These gases came from springs, seeps/vents and
tems. They were recovered by shallow coring and drilling (Lecher aquifers, located both onshore and offshore, that clearly associate
et al., 2017) or deep drilling (Wiersberg and Erzinger, 2008), and with serpentinites (Etiope et al., 2017; Nothaft et al., 2021). Some
were usually dissolved in water (Gruca-Rokosz and Koszelnik, gases in volatiles released from crushed ultramafic rocks appar­
2018). ently affected by serpentinization (Zhang et al., 2017) are also
H. Aquifers/groundwater. These gases, typically dissolved in water, included in this category.
were sampled from water wells drilled into underground layers of
water-bearing permeable rocks, rock fractures or unconsolidated There are some overlaps and uncertainties in the definitions of these
materials (Darrah et al., 2014; Goetz, 2021). This category ex­ geological habitats. For example, gas hydrates occur in marine (Milkov,
cludes gases from volcanic/magmatic hydrothermal systems, 2010) and in continental (Lu et al., 2011), including lake (Hachikubo
sedimentary hydrothermal systems, and serpentinites (see et al., 2020), sediments. However, gas hydrates contain concentrated
below). gases (Milkov, 2005), which are potentially important resources (Mil­
I. Non-sedimentary hard rocks. These gases were sampled from kov, 2004), and the process of gas hydrate formation fractionates gases
non-sedimentary igneous and metamorphic rocks (but not ser­ (Sassen et al., 2000; Milkov et al., 2004). Therefore, I differentiate gas
pentinites, see below), typically in cratonic settings (Potter et al., hydrates as a separate geological habitat. Similarly, and highly impor­
2013; Boreham et al., 2021b). The gases were obtained during tant for this study, I differentiate serpentinites as a separate habitat.
crushing of the hard rocks (Potter et al., 2013) or came from Even though gases in that category are sampled both at the surface/­
pores/cavities (Galimov and Petersil’ye, 1967; Jeffrey and mudline (seeps, springs) and in the shallow fractured aquifers, I differ­
Kaplan, 1988) and even as gas jets and seeps (Voytov, 1992). entiate that habitat strictly based on the mineralogy of the hosting rocks,
J. Volcanic/magmatic seeps. These gases were sampled at the sur­ due to the unique H2-enriched gases generated in that setting. Serpen­
face/mudline from actively (via eruptions) or passively (via fu­ tinization process can contribute to gases sampled from other habitats
maroles) degassing onshore and offshore volcanoes (excluding such as aquifers (Sherwood Lollar et al., 2007), volcanic/magmatic
vents in Mid-Ocean Ridges, see below) associated with magmatic seeps and vents in MORs (Charlou et al., 2000), sedimentary hydro­
processes (Bergfeld et al., 2014; Fischer and Chiodini, 2015; thermal systems (Hao et al., 2020), conventional reservoirs (Coveney Jr.
Fiebig et al., 2019). et al., 1987; Briere et al., 2017), and non-sedimentary hard rocks (Kelley
K. Volcanic/magmatic vents in Mid-Ocean Ridges (MORs). These and Früh-Green, 1999). However, unless the sampled gases directly
gases were sampled from vents in the Mid-Atlantic Ridge (Konn emanate from serpentinites, they have been labelled as corresponding to
et al., 2022), Mid-Cayman Rise (McDermott et al., 2018), East- other geological habitats. Gases from springs and hydrothermal systems
Pacific Rise (Proskurowski et al., 2008a, 2008b), Juan de Fuca may also be difficult to assign to one of the defined habitats. For
Ridge (Lilley et al., 1993), and the Central Indian Ridge (Kawa­ example, I classified Bazman thermal springs in the southeast Iran
gucci et al., 2010). (Deshaee et al., 2020) as “hot springs in sedimentary hydrothermal
L. Volcanic/magmatic hydrothermal systems. These gases sampled systems” even though there are nearby volcanoes with recent activities
below surface (in aquifers) were clearly related to volcano-hosted because it appears that these volcanoes contributed heat but not vola­
hydrothermal systems in which the heat and volatiles are sup­ tiles to the thermal springs. In spite of the listed limitations, the
plied from magma (Fischer and Chiodini, 2015; Fiebig et al., assignment of gases to specific geological habitats enables the holistic
2019). In the classification of Moeck (2014), these are geological and geochemical interpretation of natural gases.
convection-dominated volcanic type and plutonic type The data for gas samples include information on location, sampling,
geothermal plays exemplified, for example, by the Reykjanas depth, temperature, formation, fluid type and molecular and isotopic
system in Iceland (Arnason, 1977; Byrne et al., 2021) and the compositions. Gas concentrations are recorded as vol.% on water-free
Larderello system in Italy (Leila et al., 2021). basis. Parts of this dataset were previously described and studied by
M. Hot springs in sedimentary hydrothermal system. These gases Milkov and Etiope (2018), Milkov et al. (2020a, 2020b), Snodgrass and
sampled at the surface/mudline were apparently associated with Milkov (2020) and Milkov (2021).
deep hydrothermal systems in sedimentary rocks, with no A total of 34,290 samples in the dataset have identified geological
obvious relation to adjacent recent volcanic/magmatic activity. habitats (one of the 16 habitats discussed above, Fig. 1A). 6246 of these
Examples include hot springs in the Baikal rift zone (Kalmychkov samples have reported H2 concentrations (Fig. 1B). The concentration
et al., 2020). data come from published papers, reports and publicly available gov­
N. Sedimentary hydrothermal system. These gases came from hot ernment databases, and the validity of specific measurements is
aquifers in sedimentary rocks that do not have clear association impossible to check. Gases with reported H2 concentrations occur in 50
with volcanic/magmatic systems. In the classification of Moeck countries and territories and in MORs (Fig. 2). Most gases come from the
(2014), these are conduction-dominated geothermal plays in USA (n = 2481) and China (n = 816). A fraction of samples (n = 484)
intracratonic basins (e.g., North German basin, Feldbusch et al., have measured δ2H-H2 values.
2018), orogenic belts and basements. The extensional domain
plays with no obvious recent volcanic/magmatic activity such as
in the Western Turkey (Tut Haklıdır et al., 2021) are also included
in this category.

5
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 1. Pie charts showing the number and proportions of gases from different geological habitats in the entire dataset of gas samples with known geological habitats
(A) and in the subset of gas samples with measured H2 concentrations (B).

6
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 2. Distribution of samples with reported H2 concentrations across countries and territories (A) and across continents, including Mid-Ocean Ridges (B).

4. Results volcanic/magmatic seeps (~4%), aquifers (~3%), and conventional


petroleum reservoirs (~1%). The remaining eight geological settings
4.1. Distribution of H2 in various geological habitats and depths (marine sediments, shales, coals, hot springs in sedimentary hydro­
thermal systems, continental sediments, hydrocarbon seeps and mud
Concentrations of H2 in natural gases, when measured and reported volcanoes, and springs) have either very low average concentrations of
(n = 6246), range from 0% to 100% (Fig. 3), with average ~3.5% and H2 or/and small number of available gas samples, and I exclude them
median ~0.01%. There are 4770 gas samples with H2>0%, and this from further consideration in this study.
subset includes 1890 samples with H2>0.1%, 1012 samples with Most gas samples with significant (>10%) H2 (also referred to as H2-
H2>1%, and 621 samples with H2>5% (Fig. 3). rich by Coveney Jr. et al., 1987) occur at the surface (as seeps and vents)
The distributions of H2 concentrations in gases from 16 geological or in shallow (<2000 m) subsurface (Fig. 5). Among 1186 surface gas
settings described above are presented as box plots in Fig. 4. Gases samples with H2 measurements, the average concentration is ~10%. The
sampled in MORs and in serpentinites have the highest average con­ highest concentrations (>90%) are measured in gases seeping from
centrations of H2 (~24% and ~21%, respectively). Other six settings serpentinites (Neal and Stranger, 1983; Szponar et al., 2013) and vent­
that contain less, but still appreciable amounts of H2, include non- ing from MORs (Proskurowski et al., 2008a, 2008b; Konn et al., 2022).
sedimentary hard rocks (average ~11%), volcanic/magmatic hydro­ The average H2 in serpentinite seeps is ~23% (n = 239), and in MOR
thermal systems (~7%), sedimentary hydrothermal systems (~4%), vents it is ~24% (n = 131). In contrast, volcanic/magmatic seeps have

7
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 3. Distributions of H2 concentrations in gas samples with reported H2 concentrations. Panel A shows all available samples with bin size 0.1% and demonstrates
that the vast majority of samples has H2 concentrations <0.1%. Panel B shows all available samples with bin size 5% and also zooms into samples with H2 con­
centrations >5% (panel C) to demonstrate a relatively small number of such samples in the dataset.

significantly lower average H2 concentration of ~4% (n = 798). Samples from depth >2000 m are predominantly from conventional
At depth from >0 to 2000 m below the surface, the average H2 is reservoirs (Fig. 5). The average H2 concentration is ~1% (n = 666), but
~2.8% (n=1296 samples). The highest concentrations >90% are that value is significantly affected by few gases from deep aquifers and
measured in gases dissolved in water held by serpentinites (Sader et al., non-sedimentary rocks relatively enriched in H2.
2021; Nothaft et al., 2021), in conventional gas reservoirs (Prinzhofer It appears from the above statistics and displays in Fig. 5 that the
et al., 2018; Pudovskis, 1959) and in aquifers (Guélard et al., 2017). The average concentration of H2 is greatest at the surface and decreases with
average H2 in volcanic/magmatic hydrothermal systems is ~18.4% (n increasing depth. However, that observation results from a smaller
= 22, with highest concentrations up to 58.6% in wells from the Cerro number of samples at great depths, which come predominantly from
Prieto geothermal field, Welhan, 1981), in serpentinites ~15.5% (n = conventional reservoirs. There are no clear trends of H2 increase or
65), in non-sedimentary hard rocks ~14.4% (n = 11), in aquifers ~6.6% decrease with increasing subsurface depth in studied geological settings.
(n = 95), in sedimentary hydrothermal systems ~1.9% (n = 27), and in
conventional reservoirs ~1.2% (n = 997).

8
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 4. Concentrations of molecular H2 in natural gas samples (n=6246) from 16 geologic habitats (sorted by decreasing average values, letter labels are consistent
with text in section 3): K - Mid-Ocean Ridges; P – Serpentinites, I – Non-sedimentary hard rocks; L – Volcanic/magmatic hydrothermal systems; N – Sedimentary
hydrothermal systems; J – Volcanic/magmatic seeps; H – Aquifers/groundwaters; O – Springs; A – Conventional reservoirs; G – Continental sediments; E – Marine
sediments; F – Gas hydrates; C – Shales; B – Coals; M – Hot springs in sedimentary hydrothermal systems; D – Hydrocarbon seeps/mud volcanoes. The data are
displayed using a box plot, which shows distribution of values as histogram, average (mean) values as black stars (and 95% confidence level as black bars), median
values as dotted line, first quartile (Q1), third quartile (Q3), lower adjacent value (LAV), upper adjacent value (UAV), and outliers. The first quartile (Q1) is the
median of the lower half of the data set. This means that about 25% of the numbers in the data set lie below Q1 and about 75% lie above Q1. The third quartile (Q3) is
the median of the upper half of the data set. This means that about 75% of the numbers in the data set lie below Q3 and about 25% lie above Q3. The lower adjacent
value is the smallest observation that is greater than or equal to the lower inner fence, which is the first quartile minus 1.5×IQR, where IQR stands for the inter­
quartile range. The upper adjacent value is the largest observation that is less than or equal to the upper inner fence, which is the third quartile plus 1.5×IQR. Most
values are located in the lower part of the distributions, and displayed statistical boundaries are poorly distinguished on these box plots.

Fig. 5. Concentration of H2 (A – simple linear scale, B – logarithmic scale) at the surface and at subsurface depths in various geological habitats.

9
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

4.2. Associations of H2 with other natural gases He is generated in the crustal settings from the decay of radioactive
elements (Byrne et al., 2018), but crustal H2 either is limited at the
Hydrogen is the focus of this study, but it is, on average, not the most source or is largely consumed. On the other hand, a weak positive cor­
abundant gas even in geological habitats where it is present in relatively relation between H2 and He in gases from conventional reservoirs
high concentrations (Table 1). Other gases typically dominate, as, for potentially suggests that they are generated, migrate and/or become
example, N2 in serpentinites, CH4 in conventional reservoirs and in non- preserved in similar ways.
sedimentary hard rocks, and CO2 in volcanic/magmatic seeps and hy­
drothermal systems. Hydrogen is the second most abundant gas in MORs 4.3. Isotopic compositions of H2 and associated He
and in volcanic/magmatic hydrothermal systems. Helium is, on average,
less abundant than H2 in all these settings. Stable isotopes are a standard tool to evaluate the origin of hydro­
Fig. 6 compares concentration of H2 with the concentrations of other carbon gases, CO2, N2 and noble gases at the surface (seeps) and in the
major gases CH4, CO2 and N2 in different geological settings. Gases in subsurface (Milkov and Etiope, 2018; Lavrushin et al., 2020; Byrne et al.,
conventional reservoirs are clearly dominated by CH4, and volcanic/ 2018). Various tools ranging from two-dimensional genetic diagrams
magmatic gases are clearly dominated by CO2, and these observations (Milkov and Etiope, 2018; Milkov, 2021) to a web-based machine
are consistent with the main statistics in Table 1. There are no obvious learning tool (Snodgrass and Milkov, 2020) have been developed to
correlations between concentrations of H2 and concentrations of these interpret the origins of these gases based on large datasets of tens of
major gases in the studied geological settings. However, there is a weak thousands gas samples.
positive correlation between concentrations of H2 and He for all Geologic H2 contains two stable isotopes, 1H (protium, isotopic
considered gases (Fig. 7). Gases from MORs and volcanic/magmatic abundance ~99.972% of all hydrogen on modern Earth) and 2H
gases are more enriched in H2 and depleted in He (also see Table 1) than (deuterium). Isotopic composition of H2 is expressed as δ2H-H2 values in
gases from crustal settings. This may be because significant additional parts per thousand (per mil, ‰) referenced to Vienna Standard Mean
Ocean Water, or VSMOW (Schimmelmann and Sauer, 2017). Although
Table 1 δ2H-H2 can be indicative of H2 origin (Boreham et al., 2021a, 2021b;
Range, average and median concentrations of major gases (CH4, N2, CO2), H2 Schimmelmann and Sauer, 2017), the number of available δ2H-H2
and He in the six geological habitats with highest average H2 concentrations. measurements in natural gases is relatively small, there is a significant
Other gases that are present in smaller concentrations (e.g., ethane-pentane and overlap in δ2H-H2 values of H2 with different origins (Boreham et al.,
H2S) are not shown. The n indicates the number of samples with available 2021a, 2021b) and there is no comprehensive interpretation scheme to
concentration measurements. determine the origin of H2 based on its isotopic composition.
Geological CH4 N2 CO2 H2 He Gas Comparison of H2 concentrations with the isotopic composition of H2
habitat type suggests that there is no correlation between them for the entire set of
Mid-Ocean 0-73.3 0-88.8 0-99.9 0-97.1 0.01- CO2- gases (Fig. 8). However, there is a good separation between H2 in gases
Ridges Av. 6.7 Av. 21.6 Av. 50.6 Av. 0.05 H2- from MORs, volcanic/geothermal seeps and hydrothermal systems and
Med. 1.9 Med. 6.8 Med. 23.8 Av. N2 gases from serpentinites, conventional reservoirs and aquifers/ground­
(n = (n = 74) 47.6 Med. 0.04
waters, non-sedimentary hard rocks and sedimentary hydrothermal
137) (n = 4.1 Med.
112) (n = 0.04 system based on δ2H-H2 of -650‰.
137) (n = The distribution of δ2H-H2 values across various geological settings is
12) broadly trimodal (Fig. 9). Molecular H2 from serpentinites (Fig. 9b),
Serpentinites 0-95.1 0-99.5 0-99.9 0-100 0-0.01 N2- conventional petroleum reservoirs, aquifers, non-sedimentary hard
Av. 26.5 Av. 36.1 Av. 16.8 Av. Av. CH4-
Med. 9.7 Med. Med. 21.2 0.65 H2
rocks and sedimentary hydrothermal systems has δ2H-H2 values gener­
(n = 25.0 (n 0.05 (n Med. Med. ally more negative than -650‰ (mode around -730‰). Not included in
428) = 390) = 266) 4.8 0.01 Fig. 9 are gases extracted from shales and metapelites with δ2H-H2
(n = (n = values from -810 to -629‰ (Suzuki et al., 2017). In contrast, molecular
397) 248)
H2 from volcanic/magmatic seeps and hydrothermal systems is rela­
Non- 0.1-92.9 0-99.7 0-92.8 0-63.3 0-19.1 CH4-
sedimentary Av. 48.1 Av. 36.8 Av. 3.5 Av. Av. N2- tively enriched in 2H with δ2H-H2 values generally exceeding -650‰
hard rocks Med. Med. Med. 0.2 8.1 3.0 H2 (Fig. 9c), with apparent two modes around -530‰ and -370‰. Gases
52.0 32.4 (n = 70) Med. Med. sampled from vents in MORs are generally even more enriched in 2H
(n = 86) (n = 83) 3.0 (n 1.1 with δ2H-H2 values with a mode about -370‰ (Fig. 9d).
= 77) (n =
80)
These observations suggest that δ2H-H2 values may be indicative of
Volcanic/ 0-74.3 0-97.6 0-99.6 0-64.7 0-19.9 CO2- H2 origin: hydrogen with δ2H-H2 of <-650‰ has crustal origin, while
magmatic Av. 5.9 Av. 6.7 Av. 79.9 Av. Av. H2- hydrogen with δ2H-H2 of >-650‰ has mantle/primordial origin. Pri­
hydrothermal Med. 0.9 Med. 0.9 Med. 6.8 0.9 N2 mordial H2 was delivered by meteorites during accretion of the Earth
systems (n = (n = 84.1 (n Med. Med.
and preserved in the mantle and the core (Larin, 1993, 1995). The
156) 151) = 137) 1.6 (n 0.0
= (n = separation of crust-sourced and mantle-sourced gases based on δ2H-H2
149) 74) of -650‰ is supported by the well-established source fingerprinting
Volcanic/ 0-96.1 0-99.0 0-99.9 0-80 0-33.3 CO2- based on He isotopes as shown in Fig. 10 (3He/4He ratios are typically
magmatic Av. 2.1 Av. 10.5 Av. 81.4 Av. Av. N2- reported relative to the atmospheric ratio Ra and expressed as R/Ra).
seeps Med. 0.2 Med. 1.6 Med. 2.6 0.02 H2
(n = (n = 91.8 (n Med. Med.
Helium from the mantle is relatively enriched in 3He, which is a pri­
848) 817) = 803) 0.2 (n 0.00 mordial isotope made in the Big Bang and incorporated into Earth
= (n = during its initial accretion (Byrne et al., 2018). The association of 2H-
767) 640) enriched H2 with 3He-enriched He suggests that 2H-enriched H2
Conventional 0-100 0-97.8 0-100 0-97.4 0-9.0 CH4-
observed in volcanic/magmatic natural gases has mantle and/or pri­
reservoirs Av. 81.6 Av. 5.3 Av. 4.7 Av. Av. N2-
Med. Med. 1.7 Med. 1.0 0.9 0.15 CO2 mordial origin. Gas samples from the crustal geological settings are
87.1 (n (n = (n = Med. Med. relatively depleted in 3He and in 2H (Fig. 10). Specific sources of He and
= 10,251) 12,366) 0.0 (n 0.02 H2 in each gas sample from crustal settings can be different (e.g., He can
14,660) = (n = come from the decay of radioactive elements and from the air, while H2
2150) 2688)
is sourced by the reactions of serpentinization, Vacquand et al., 2018) or

10
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 6. Concentration of H2 versus concentrations of CH4, CO2 and N2 in natural gases from various geological habitats. See Fig. 5 for symbol legend.

Fig. 7. Concentration of H2 versus concentrations of He in natural gases from various geological habitats. See Fig. 5 for symbol legend. Samples that contain one gas
(for example, H2), but not the other (respectively, He), as, for example, some gases emanating from serpentinites in Oman (Zgonnik et al., 2019) are not displayed on
this log-log plot.

the same (Naumenko-Dézes et al., 2022). as high as -110‰. This H2 is associated with mantle/primordial He, and
Gases venting at MORs have a wide range of δ2H-H2 values from -823 may have mantle/primordial origin. Among these samples, δ2H-H2
to -231‰ (n = 52). Three samples from the East Pacific Rise extremely values closer to -700‰ may indicate H2 of mixed mantle/crustal origin,
depleted in 2H (δ2H-H2 values around -800‰, Fig. 9d) are likely sam­ for example in some samples from the Yellowstone (Welhan, 1981). δ2H-
pling artifacts (Merlivat et al., 1987). Ten samples with δ2H-H2 values H2 values closer to -400‰ (as in MOR samples, Fig. 9) and to -100‰ (as
between around -700 and -600‰ come mostly from the Lost City, where in the Mid-Icelandic volcanic belt, Sigvaldson and Elisson, 1968; Arna­
the process of serpentinization is very well documented (Lang and son and Sigurgeirsson, 1968) may indicate H2 from processes in the
Brazelton, 2020). The vast majority of MOR samples have δ2H-H2 values mantle, and perhaps some H2 of primordial origin. Such interpretation is
between -500‰ and -230‰ (Fig. 9d), suggesting mostly mantle/pri­ consistent with δ2H-H2 values mostly between -95‰ and -35‰ in min­
mordial origin. erals and fluid inclusions in mantle-derived rocks (Kyser and O’Neil,
The isotopic composition of molecular H2 in volcanic and magmatic 1984; Deloule et al., 1991; Kingsley et al., 2002; Loewen et al., 2019),
settings is distinct, with δ2H-H2 values generally greater than -700‰ and the suggestion that primordial H2 has δ2H-H2 of about -75‰ (Craig and

11
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 8. Concentration of H2 versus δ2H-H2 in various geological habitats. See Fig. 5 for symbol legend. The value of δ2H-H2 -650‰ may separate crustal and mantle/
primordial H2 (see text for discussion).

Lupton, 1976; Taylor et al., 1983; Larin, 1995) and even greater 2H genetic diagrams with plotted gases that contain ≥1% of H2. It appears
enrichment in asteroids and comets (Dauphas et al., 2000). that hydrocarbons and CO2 sampled from conventional reservoirs have
The variations of δ2H-H2 in geological habitats are interpreted above mostly thermogenic origin. Hydrocarbons and CO2 sampled from other
in terms of different sources and origins of H2 (also complicated by the geological settings, especially volcanic gases, gases from serpentinites
fact that the geological habitats do not exactly correspond to specific and gases from non-sedimentary hard rocks have mostly abiotic origin.
sources of hydrogen). Variations in source water isotope composition However, it is important to note that the construction of gas genetic
(δ2H-H2O) and temperature-dependent equilibrium isotopic fraction­ fields on these diagrams was grounded in both geochemical and
ation and kinetic 2H-1H exchanges between H2 and water (Pester et al., geological arguments (Milkov and Etiope, 2018), so the conclusions
2018; Ricci et al., 2022; Löffler et al., 2022) provide additional reasons above are based on circular reasoning. In addition, many natural gases
for these variations or an alternative explanation of such variations. The are mixtures. Almost no gases that contain ≥1% of H2 have hydrocar­
fractionation factor at equilibrium (αH2O-H2(eq)) is sensitive to tempera­ bons and CO2 of microbial origin, which may be explained by the close
ture (Richet et al., 1977; Horibe and Craig, 1995). In general, H2 association of H2 producers and consumers in shallow cool marine
sampled at high temperatures and/or sourced by high-temperature sediments and oil pools where microbial gases typically form.
fluids should be more enriched in 2H (i.e., should have more positive The concentration of H2 has a weak correlation with the values of
δ2H-H2 values) than H2 sampled at low temperatures. Data from natural δ13C-CH4 (Fig. 14). Methane in gas samples (all geological settings) with
gases generally follow this trend (see Fig. 1 in Pester et al., 2018), but >10% H2 has average δ13C value -21.8‰ (n = 172), while methane in
there are many deviations from the equilibrium curve, which can be gas samples with H2 content >0% and ≤10% has average δ13C value
explained by previous higher temperatures of the solutions. There is a -38.0‰ (n = 1509). The presence of very high H2 concentrations
weak positive correlation between δ2H-H2 values and sampling tem­ together with 13C-enriched methane likely indicates common abiotic
perature in our dataset (Fig. 11), which is broadly consistent with processes in which these gases are co-generated (Etiope and Sherwood
temperature-dependent isotopic fractionation between H2 and water. Lollar, 2013; Etiope and Schoell, 2014; Etiope and Whiticar, 2019).
Comparison of δ2H-H2 values with carbon and hydrogen isotopes of
co-existing CH4 suggests that crustal and mantle CH4 can be distin­
4.4. Origins of H2-containing gases guished based on δ2H-CH4, with a separation at about -200‰ (Fig. 15).
While carbon isotopes of CH4 help distinguish between biotic and
Molecular H2 is commonly a relatively minor compound in natural abiotic processes, the plot of δ2H-H2 versus δ13C-CH4 highlights that
gases dominated by hydrocarbons (predominantly CH4), CO2 and N2 gases from crustal geological settings can have biotic or abiotic origins,
(Table 1). The origins of these major gases can be inferred from stable while mantle/primordial gases have abiotic origin.
isotopes (Milkov and Etiope, 2018; Snodgrass and Milkov, 2020; Milkov,
2021; Lavrushin et al., 2020). Figs. 12 and 13 show commonly used

12
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 9. Histograms showing distributions of δ2H-H2 values in gas samples from eight geological settings (A), and separately from serpentinites (B), volcanic/
magmatic settings (C) and MORs (D). The bin size is 20‰.

4.5. H2 in conventional reservoirs δ2H-H2 value in the Heins and Scott gases is -753‰ (n = 11), which is
close to isotopic composition of H2 sourced by serpentinites (Fig. 9b).
Conventional reservoirs with reported H2 concentrations are the Gas with high (98%) concentration of H2 (Fig. 5) was serendipitously
most represented geological settings with the largest number of H2 discovered in the Bourakebougou area (Mali) by a water well (Briere
measurements in the studied dataset (n = 2237). The average concen­ et al., 2017; Prinzhofer et al., 2018). The initial study speculated that
tration of H2 in gases from conventional reservoirs is ~0.8%, while the serpentinization of ultramafic rocks is one of the most likely mechanisms
median concentration is only 0.01% (Fig. 4). I note that these statistics of H2 generation in the area (Briere et al., 2017). The later, more detailed
are for all samples in the dataset, and they are not weighted for gas integrated geological/geochemical study suggested that H2 (it has δ2H-
resources in conventional reservoirs. The average concentration value is H2 value -702‰) comes from the basement, without specifying the
likely to be significantly pushed upwards due to 38 samples with >10% process of H2 generation (Prinzhofer et al., 2018). Serpentinized mantle
of H2 from several poorly understood gas accumulations in Australia and peridotites are present in the NE Mali (the Timétrine massif, Caby, 2014)
better studied accumulations in Kansas and in Mali. but the presence of serpentinites in the proximity of the Bourakebougou
High concentrations of H2 in conventional reservoirs in Australia area located within the West African craton in SE Mali is not apparent
(Fig. 5) have been reported long time ago, for example 51-84% in the from the literature review.
Adelaide area within the Gawler Craton (Ward, 1933) and up to 95.3% Curiously, very high concentration of H2 of 88.5% (with δ2H-H2
in the Meda field in the Canning Basin (Pudovskis, 1959). These high H2 -742‰) was reported in an oil-associated gas from the Brent field (211/
concentrations were highlighted in the recent review by Boreham et al. 29-B17) in the North Sea (UK OGA, 2021). That gas was unusual because
(2021a), but they should be validated with modern measurements most gases from the Brent field do not contain H2. The high H2 was likely
(including isotopes) to be better understood. formed via artificial processes such as metal corrosion. Apparently
High concentrations of H2 were observed in the Heins and Scott wells artificial gases with high H2 (up to 90%) found in a well annuli on Troll A
in Kansas (Fig. 5) where 31 gas samples have H2 concentrations ranging platform offshore Norway have δ2H-H2 values about -750 to -725‰
from 1.4 to 70% (average ~32%), the remainder being N2 and traces of (Thiessen et al., 2009). The δ2H-H2 values of artificial H2 may overlap
hydrocarbons (McCarthy Jr. et al., 1986; Goebel et al., 1984; Coveney with the δ2H-H2 values of H2 from serpentinization (Fig. 9b), making
Jr. et al., 1987; Guélard et al., 2017). The Sue Duroche #2 water well separation of these processes based on H2 isotopes alone rather
drilled in the same area also contains high concentrations of H2 (Guélard challenging.
et al., 2017). These wells are near a major rift system with mafic rocks Plotting of δ2H-H2 values versus concentrations of H2 in conventional
and known kimberlites, which led to the initial interpretation that H2 reservoirs (Fig. 16) reveals that: 1) there is a wide range of δ2H-H2 values
originated from the reactions of serpentinization (Coveney Jr. et al., from -996‰ to -370‰ (n = 74), but most samples (n = 69) have δ2H-H2
1987), although the later study presented a more complex interpretation values from -850‰ to -600‰; 2) there is essentially no correlation of
of both deep and shallow H2 origins (Guélard et al., 2017). The average δ2H-H2 values and concentrations of H2. Three samples more depleted in

13
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 10. Isotopic composition of He (expressed as R/


Ra, see Byrne et al., 2018) versus isotopic composi­
tion of H2 (δ2H-H2) in gas samples from eight
geological settings. Three sets of samples (labeled)
are displayed but excluded from interpretation:
samples from 13N East Pacific Rise (Merlivat et al.,
1987) and from Groß Schönebeck (Feldbusch et al.,
2018) were interpreted as artifacts in the original
studies, and one sample from the Semail ophiolite
(Sano et al., 1993) is an outlier from other samples
taken in the same area. The other samples are inter­
preted as groups of mantle and primordial He and H2
and as crustal He and H2. Still, δ2H-H2 values may
reflect temperature-dependent equilibrium isotopic
fractionation and kinetic exchanges between H2 and
water (see the text for that discussion).

2
H (δ2H-H2 <-850‰) than most other samples are from the West Other gases with δ2H values from -850 to -600‰ displayed in Fig. 16
Moonfish, Marlin and Basket wells in the Gippsland basin offshore come from fields (e.g., Brent field in the North Sea) and basins (Bohai
Australia (Boreham et al., 2021a). Such δ2H-H2 values are highly un­ Bay in China, Taoudeni in Mali, and Gippsland, Bonaparte, Canning and
usual in the studied dataset (Fig. 9a). There are two samples more Bass basins in Australia) that do not have known nearby occurrences of
enriched in 2H (δ2H-H2 >-600‰) than most other samples (Fig. 16), serpentinites. Other crustal sources and H2 generation mechanisms may
including one sample from Australia (the Katandra well in the Bonaparte be responsible for the presence of H2 in these gases. δ2H-H2 values in
basin). Boreham et al. (2021a) suspected that artifacts accompanying conventional reservoirs are similar to δ2H-H2 values (from -810 to
gas production, collection or storage caused unusual δ2H-H2 values in -629‰) in H2 released from pulverized shales and metapelites and
these gases from Australia. The fifth unusual sample is from the Heins supposedly liberated from the organic matter (Suzuki et al., 2017). This
well in Kansas, but that single value δ2H-H2 >-520‰ reported by suggests that H2 in conventional reservoirs may come from the organic
McCarthy Jr. et al. (1986) is outside of the range of the other 16 mea­ matter (Horsfield et al., 2022) and perhaps from the same source rocks
surements from -670‰ to -836‰ (average -779‰) in the Heins and that supplied the major gases such as hydrocarbons and CO2. However,
Scott wells (Goebel et al., 1984; Coveney Jr. et al., 1987; Guélard et al., it is not clear at this stage if other most often mentioned natural pro­
2017), making that value suspicious. cesses (e.g., radiolysis or water or fracture-induced reduction of water)
Samples with a significant 2H enrichment may be contaminated with or some other processes listed in Section 2.2. contribute H2 to conven­
H2 from corrosion of steel (Levshounova, 1991) or may contain abiotic tional reservoirs. A much larger dataset of δ2H-H2 values integrated with
mantle-derived H2 from some undelaying deep-seated rocks (Smith the geological settings of the conventional wells is needed to answer this
et al., 2005; Truche and Bazarkina, 2019). Still, gases in conventional question.
reservoirs have δ2H-H2 values mostly from -850‰ to -600‰ (Fig. 16),
suggesting that these processes may be not important in conventional 5. Discussion
reservoirs.
Similarity of δ2H-H2 values between conventional reservoirs and 5.1. The value and use of δ2H measurements in the interpretation of H2
serpentinites (Figs. 9 and 16) suggests that reactions of serpentinization origins
may be responsible for the origin of H2 in some conventional reservoirs.
This is likely the case for the H2-rich gas in the Heins and Scott wells in Although δ2H-H2 can be indicative of H2 origin (Schimmelmann and
Kansas. Interestingly, methane in these wells has δ2H values from -477 to Sauer, 2017; Boreham et al., 2021a, 2021b), previous studies evaluated
-419‰ (n = 5) and plot on the genetic diagram of δ13C-CH4 versus δ2H- relatively small datasets of available δ2H-H2 measurements in natural
CH4 in the area of abiotic gas near many samples from serpentinites gases with significant overlaps in δ2H-H2 values of H2 with different
(Fig. 12). This methane may have formed via FTT reactions from the H2 origins (Hao et al., 2020; Boreham et al., 2021a, 2021b) and did not
that originated from serpentinization of mafic rocks and kimberlites propose a comprehensive interpretation scheme to determine the origin
documented in the area (Coveney Jr. et al., 1987). of H2 based on its isotopic composition.

14
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 11. Sampling temperature versus δ2H-H2 values in natural gases from different geological settings. See Fig. 5 for symbol legend. The dashed line represents
calculated δ2H-H2 values assuming δ2H-H2O=0‰ for equation α[H2O-H2] = 1.0473+201036/T2 + 2.060×109/T4 + 0.180×1015/T6, which describes temperature-
dependent (temperature is in Kelvin) hydrogen isotopic equilibrium (exchange) between H2 and H2O (Horibe and Craig, 1995; Kawagucci et al., 2010).

Fig. 12. Gas genetic diagrams (from Milkov and Etiope, 2018) with displayed gas samples that contain H2 concentrations ≥1%. See Fig. 5 for symbol legend. The
ratio of C1/(C2+C3) is the ration of methane to the the sum of ethane and propane, also known as “dryness” of the gas. Abbreviations: CR – CO2 reduction, F – methyl-
type fermentation, SM – secondary microbial, EMT – early mature thermogenic gas, OA – oil-associated (mid0mature) thermogenic gas, LMT – late mature ther­
mogenic gas.

The results of this study suggest that crustal and mantle/primordial isotopic fractionation and kinetic isotope exchanges between H2 and
molecular H2 have distinctive δ2H-H2 values: crustal H2 is relatively water (Horibe and Craig, 1995; Kawagucci et al., 2010; Simon et al.,
more depleted in 2H (δ2H-H2<-650‰) and mantle/primordial H2 is 2011; Pester et al., 2018; Ricci et al., 2022). Many δ2H-H2 values in
relatively more enriched in 2H (δ2H-H2>-650‰). Further identification natural gases plotted against temperature generally follow the theoret­
of specific sources of H2 based on δ2H may be possible as discussed ical trend calculated for temperature-dependent hydrogen isotopic
below. Still, I caution that this discussion should be considered as pre­ equilibrium (exchange) between H2 and H2O (Fig. 11). Most δ2H-H2
liminary and subject to significant future revisions because the number values that deviate from that trend may be explained by fluid cooling
of currently available δ2H-H2 measurements is relatively small. Another and the subsequent isotope equilibrium reaction (Kawagucci et al.,
important consideration that should be integrated into the interpreta­ 2010) and the variations in source water isotope composition (δ2H-
tion of δ2H-H2 values in natural gases is the variations in source water H2O). It is possible that isotope exchange (equilibrium) reaction is a
isotope composition (δ2H-H2O) and temperature-dependent equilibrium dominant process that controls δ2H-H2 values in natural gases. The

15
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

(along with He and Ne) has been proposed a long time ago (Chamberlin,
1909; Craig and Lupton, 1976). The mantle/primordial H2 is docu­
mented by measurements in volcanic and magmatic gases (Fischer and
Chiodini, 2015), is theoretically feasible, and may be quantitatively very
significant (Gilat and Vol, 2005, 2012; Yang et al., 2016; Mao et al.,
2017; Komabayashi, 2021). The contribution of mantle-sourced H2 to
shallow gas accumulations has been previously discussed (Meng et al.,
2015). The amount of H2 in the Earth is debatable, but some studies
suggest that the interior of the Earth is enriched in H2 (Larin, 1993;
Walshe, 2006; Ikuta et al., 2019; He et al., 2022). Larin (1993) and
Toulhoat and Zgonnik (2022) estimated that the early Earth contained 4
wt.% initial H2. This H2 stored in hydrides could progressively decom­
pose and degass from the surface over geological time (Larin, 1993). The
H2 degassing was more significant for early Earth. It was proposed that
an Archean atmosphere contained up to 30% H2 (Wordsworth and
Pierrehumbert, 2013) and that gases in Archean igneous rocks contain
significantly more H2 than gases from more recent igneous rocks
(Chamberlin, 1909). Still, the importance of mantle and primordial H2 in
natural gases remains speculative (Zgonnik, 2020) and this origin
received no serious consideration in recent studies and reviews of H2
Fig. 13. Gas genetic diagram (from Milkov, 2021) with displayed gas samples origins on Earth (Klein et al., 2020). Our results suggest that primordial
that contain H2 concentrations ≥1%. See Fig. 5 for symbol legend. H2 may vent at the surface of the Earth in MORs and in some volcanic
settings (Sigvaldson and Elisson, 1968; Kiyosu, 1983; Lyon and Hulston,
broad correlation of δ2H-H2 values with temperature (Fig. 11) does not 1984). Primordial H2 may support unique ecosystems with chemo­
directly contradict the variation of δ2H-H2 values due to origins from lithoautotrophic metabolic pathways on the modern deep seafloor
different geological settings. High-temperature geological settings such (Takai et al., 2004; Ohara et al., 2012), provide energy for long-term
as volcanoes may emit H2 that is relatively enriched in 2H because it has survival of microbial cells (Morita, 2000) and contribute to the origin
primordial/mantle origin or/and because it is at equilibrium with water of life on Earth and perhaps other planets (Nisbet and Sleep, 2001;
vapor at that high temperature. McCollom and Seewald, 2013; Holm et al., 2015; Glein and Zolotov,
Molecular H2 most enriched in 2H (δ2H-H2>-350‰) perhaps is 2020).
mostly primordial. Degassing of primordial H2 at the Earth surface Molecular H2 from volcanic/magmatic settings has a bimodal

Fig. 14. Concentration of H2 versus δ13C-CH4 in various geological habitats. See Fig. 5 for symbol legend.

16
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Fig. 15. δ2H-H2versus δ13C-CH4 (A) and δ2H-H2 versus δ2H-CH4 (B) in various geological habitats. Broadly defined fields of “crustal” and “mantle/primordial” gases
are outlined. See Fig. 5 for symbol legend.

Fig. 16. Concentration of H2 versus δ2H-H2 values in conventional petroleum accumulations and in serpentinites.

17
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

distribution of δ2H values, with one mode around -380‰ (similar to the H2 concentrations in the SAFOD well. In one experiment of simulated
MORs) and the second mode around -520‰ (Fig. 9c). This may indicate faulting of a basalt specimen dampened with water with δ2H -40‰, the
that some volcanoes emit mostly primordial 2H-enriched H2, maybe δ2H-H2 value was -222‰ (Hirose et al., 2011). Mechanoradical H2 from
from the deeper mantle and from the core, for example, in the Mid- the single lab-simulating faulting experiment is significantly enriched in
2
Icelandic volcanic belt (Sigvaldson and Elisson, 1968; Arnason and H relative to the geologic H2 in fault settings. There may be many
Sigurgeirsson, 1968). Other volcanoes emit relatively more 2H-depleted reasons for that, including mixing of H2 with different origins in the
H2 generated in the upper mantle or mixed with some crustal-generated subsurface (e.g., H2 from serpentinites, from the mantle, or from other
H2 (Welhan, 1981). subsurface processes may mix with H2 generated at surfaces of fractured
Serpentinization is one of the main processes that can result in high rocks), isotopic fractionation of mechanoradical H2 by H2-metabolizing
H2 concentration in crustal geological settings (Coveney Jr. et al., 1987; organisms around natural fault zones, and laboratory experiments
Etiope et al., 2011), and our results support this view. Gases sampled poorly imitating natural conditions. The quantitative significance of the
from serpentinites contain molecular H2 with δ2H values around -715‰ reaction between groundwater and Si-O or Si- radicals in the generation
(Fig. 9b). Such 2H-depleted H2 may be due to specific serpentinization of subsurface H2 remains unclear and more field and laboratory mea­
reactions or just a function of relatively low temperatures at which H2 surements of H2 concentrations and δ2H-H2 values are needed to
equilibrates with water in serpentinite settings (Fig. 11). constrain it.
Suzuki et al. (2017) report that H2 released from pulverized shales The molecular H2 generated by microbes through hydrogenase re­
and metapelites sampled in one well and in outcrops in Japan has δ2H action in laboratory conditions has δ2H-H2 values between -814‰ and
values from -810 to -629‰ (n = 15, average -696‰). They interpreted -763‰ (Krichevsky et al., 1961). However, H2-enriched natural gases
that this H2 was liberated from the organic matter (kerogen) during usually do not have microbial origin (Fig. 12). This may be because in
thermal maturation. These δ2H values are similar to δ2H values in gases geological settings favorable for the generation of microbial gases (e.g.,
sampled from conventional reservoirs (n = 72, average -722‰, Fig. 9). shallow cool marine sediments), the H2 is rapidly consumed and
Although significant amounts of H2 were obtained during laboratory metabolized into hydrocarbons, mostly methane (Wagner et al., 2018;
pyrolysis of shale and coals (e.g., Li et al., 2015, 2017; Horsfield et al., Katz, 2011; Milkov, 2011).
2022), there are no isotope measurements on pyrolysis-derived H2 (to Artificial H2 present in some sampled natural gases may complicate
our knowledge). the use of δ2H measurements in the interpretation of H2 origin. Artificial
Radiolysis of water by alpha, beta and gamma radiation released in H2 may be generated in the subsurface during drilling operations via
the radioactive decay of U-, Th- and K-bearing minerals (Boyd et al., reactions between acids introduced into the wellbore or naturally pre­
1973; Lin et al., 2005a, 2005b; Sauvage et al., 2021) and radiolytic sent in groundwaters and the iron in well casings, steel drilling shavings
dehydrogenation of kerogen and hydrocarbon molecules (Frolov et al., and lithic fragments (Zinger, 1962; Bjornstad et al., 1994; Feldbusch
1998; Larter et al., 2019; Wang et al., 2022) are among other commonly et al., 2018; Halas et al., 2021), as experimentally reproduced by
discussed process that can explain the presence of H2 in the crustal gases Bjornstad et al. (1994). The corrosion process have been invoked to
(Leila et al., 2022). Laboratory experiments indicate that gamma ray explain unusually high concentrations of H2 in several wells. For
irradiation of water results in δ2H-H2 values from -348 to -539‰ (Lin example, Lyon and Hulston (1984) interpreted H2 in one geothermal gas
et al., 2005a, 2005b) and gamma ray radiolysis of methane results in sample with anomalous δ2H-H2 values around -600‰ (compared to δ2H-
δ2H-H2 values around -400‰ (Boreham and Davies, 2020). Molecular H2 values of >-525‰ measured for the rest of the samples in their study)
H2 with such δ2H values is not present in crustal gas samples investi­ to result from many minor effects such as H2S reduction to H2 by reac­
gated in this study. This may be because of relatively rapid (on tion with the steel well casing. Unusually high H2 (9.2%) with δ2H value
geological timescales) isotopic exchange between H2 and groundwater -277‰ in a gas sample from geothermal reservoir Groß Schönebeck
(Lin et al., 2005a, 2005b; Ricci et al., 2022). Alternatively, radiolysis of (North German Basin) was judged to result from corrosion during a long
water and hydrocarbons is not a quantitatively significant process of H2 exposure of steel well casing at pH 2 (Feldbusch et al., 2018).
generation in the crust. Experimental data from Wang et al. (2020) show Merlivat et al. (1987) observed unusual δ2H-H2 values from -823‰ to
that H2 generated from irradiated kerogen (δ2H-H2 value -170‰) has -759‰ in H2-rich samples taken above hydrothermal vents in the East
δ2H-H2 values from -710‰ to -450‰. That range of δ2H-H2 value Pacific Rise and interpreted such H2 as a result of a reaction between the
partially overlaps with δ2H-H2 values in gases from conventional res­ titanium sample bottles and the hot and acid hydrothermal solutions.
ervoirs (Fig. 9). This suggests that radiolysis of kerogen may produce at Such δ2H-H2 values are indeed anomalous among the larger dataset of
least some H2 observed in conventional reservoirs. samples from MORs, where H2 commonly has δ2H values >-500‰
Reaction between groundwater and Si-O or Si- radicals that appear (Fig. 9d).
on the fresh surfaces of fractured rocks in fault zones is another often- The δ2H-H2 values of samples with high (up to 90%) and apparently
implied mechanism of H2 generation in the subsurface (mechanor­ artificial (due to CO2 corrosion) H2 in well annuli on Troll A platform
adical H2, Kita et al., 1982). Elevated concentrations of H2 often (North Sea, Norway) were mostly about -750 to -725‰ (Thiessen et al.,
observed in soil gases sampled along fault zones support this hypothesis 2009; Mørkved et al., 2009). Using the H2O(LIQ)-H2 geo-thermometer
(Wakita et al., 1980; Sugisaki et al., 1983). Soil gases sampled from sites the predicted equilibrium temperatures were, with some exemptions,
along the Yamasaki fault (Japan) displayed molecular H2 concentrations in the temperature range of the Troll A well heads (15-50◦ C), where the
in the 0.7% to 3% range with δ2H-H2 values -770‰, -590‰, -510‰ and H2 likely formed. A sample with likely artificial H2 concentration 88.5%
-470‰ (Kita et al., 1980). Sugisaki et al. (1983) reported δ2H-H2 value from the Brent B platform has δ2H-H2 value -742‰ (UK OGA, 2021).
-675‰ in soil gas in the area of the Atera fault (Japan). High concen­ The preferential leakage of light isotope 1H during long-term
trations of H2 (up to 25%) were observed in gases extracted from a drill (months and years) storage of gas samples in steel cylinders may lead
core taken in the fracture zone in the Nojima Fault Zone in Japan (Arai to erroneously high δ2H-H2 values. This was noted, for example, for
et al., 2001), but there were no isotope measurements. High concen­ some of the first samples collected at Heins and Scott H2-rich wells in
trations of H2 (up to 6%) with δ2H-H2 values from -708‰ to -618‰ were Kansas that had δ2H-H2 values from -510‰ to -340‰ after being stored
observed in drill-mud gases sampled in fractured zones around the San for >2 months (Goebel et al., 1984). These samples were significantly
Andreas Fault core penetrated by the SAFOD well in California (Wiers­ more enriched in 2H than the H2 sampled in later studies and converted
berg and Erzinger, 2008). The authors found that high H2 is consistent to water right in the field (Coveney Jr. et al., 1987). Boreham et al.
with interaction of water with fresh silica mineral surfaces in the fault, (2021a) suspected that extreme enrichment of 2H (relative to other
but noted that serpentinites were observed in drill-cuttings (Wiersberg gasses) in two gases in their dataset (from well Pinelands 3 (Surat Basin)
and Erzinger, 2008), and reactions of serpentinization could cause high with δ2H-H2 value -389% and from well Katandra 1 (Bonaparte Basin)

18
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

with δ2H-H2 value -370%, Fig. 8) point to the artifact accompanying gas accepted because W.E. Pratt “could not point to a quantitatively
production, collection or storage. adequate natural source of free hydrogen to account for the [hydroge­
The δ2H-H2 values of artificial H2 (from -823‰ to -277‰, based on nation] process” (Hawkes, 1972, p. 2268). Hawkes (1972) listed some
the above references) and of leaked gas samples (significantly 2H- natural occurrences with high concentrations of molecular H2 that
enriched) may overlap with the δ2H values of H2 in samples where these became known since Pratt’s publication, but admitted that reports of H2
issues are not suspected. This indicates a significant challenge for the in the subsurface are subject to review due to extreme difficulties with
interpretation of δ2H-H2 values and highlights the need for careful the reliable collection, transportation and analysis of gas samples with
sampling procedures and a holistic approach to the interpretation of H2 significant H2.
origin when some geochemical anomalies or unusual gas sampling The conventional theory of hydrocarbon formation in the subsurface
conditions are present. implies that organic matter in the source rocks supplies hydrogen to
form hydrocarbons (Tissot and Welte, 1984; Hunt, 1996), while the
5.2. H2 in petroleum fluids influence of exogenous H2 on hydrocarbon generation is given little
consideration (Wang et al., 2018). It appears from this review that there
There are some petroleum (mostly natural gas) accumulations where are various sources of H2 in sedimentary basins and petroleum systems.
H2 is reportedly present in relatively large concentrations exceeding Laboratory studies suggest that large amounts of H2 can be released from
10% (Bohdanowicz, 1934; Headlee, 1962; Molchanov, 1981; Angino the radiolysis of waters and from irradiation of organic matter in
et al., 1984; Perevozchikov, 2011; Leila et al., 2022). Large concentra­ uranium-rich shales (and other rocks). That H2, if indeed generated in
tions (up to 43%) of H2 were found in many water samples taken below the subsurface in significant amounts, can react with undersaturated,
oil accumulations in Lower Volga region in Russia (Zinger, 1962). A well long-chain hydrocarbons as well as sulfur and nitrogen compounds to
targeting a petroleum prospect in Kazakhstan encountered a free water produce saturated and short-chain hydrocarbons per following reactions
flow with 90-98% of molecular H2 in dissolved gas (Kraychuk, 1976). (Wang et al., 2019):
More recently, Prinzhofer et al. (2018) reported a unique serendipitous R’CH = CH2 + H2 (catalyst) → R’CH2CH3
discovery of a large free gas accumulation that contains 98% H2, 1% N2 RCH2CH2CH2R’ + H2 → RCH2CH3 + R’CH3
and 1% of thermogenic hydrocarbon gases (mostly methane) in Prote­ R-SH + H2 → RH + H2S
rozoic sedimentary rocks in Mali. Some researchers compared that dis­ Results from a laboratory experiment by Wang et al. (2006) sug­
covery to the pioneering Drake’s well that set off the first oil rush in gested that water provided exogenous H2 for hydrocarbon generation
1859 pointing out that the well in Mali can start the rush for H2 from lignite. Wang et al. (2018) discussed a significant possibility that
exploration (Gaucher, 2020). However, the fundamental understanding uranium abundant in the Chang 7 Member source rock in the Ordos
of a connection between oil and gas and molecular H2 is lacking (Lev­ Basin (China) radiated formation water and produced H2 for hydrocar­
shounova, 1991; Smith et al., 2005). bon generation by that source rock.
Near-surface fermentation by microorganisms is the only biotic Szatmari (1989) proposed that petroleum liquids can form by
process that is considered among main sources of H2 in the Earth’s crust Fischer-Tropsch synthesis on magnetite and hematite catalysts when
(Truche et al., 2020). Virtually all of that H2 is converted to methane and CO2 from metamorphic/igneous decarbonation of carbonates reacts
the rest escapes into the atmosphere (Conrad, 1995). In deeper sedi­ with H2 derived from serpentinization of shallow-mantle lithosphere
ments, hydrogenotrophic microbes (Wagner et al., 2018) utilize H2 and and ophiolite thrust sheets. The Fischer-Tropsch synthesis of natural
CO2 to produce primary microbial (e.g., Katz, 2011) and secondary gases is supported by field observations and geochemical studied in
microbial (Milkov, 2011) gases (mostly methane), which together ac­ areas with serpentinites (Etiope et al., 2011, 2016; Etiope and Whiticar,
count for 8-15% of the global endowment of recoverable gas in con­ 2019). Mostly gaseous hydrocarbons were produced during abiotic
ventional petroleum reservoirs (Milkov, 2018). Levshounova (1991) synthesis under hydrothermal conditions in laboratory experiments
used observations from shallow sediments to propose that H2 can hy­ (McCollom and Seewald, 2006). Natural hydrocarbon gases clearly are
drate olefins (ethylene, propylene and buthylenes) resulting in the for­ generated via the Fischer-Tropsch reactions in serpentinites settings, but
mation of normal alkanes (ethane, propane, butanes) but that the proportion of such gases in the global endowment of hydrocarbon
hypothesis has not been sufficiently tested yet. gases (Milkov, 2018) is likely very small. Further, the existence of liquid
While the role of molecular H2 in making natural microbial gases (i. petroleum from Fischer-Tropsch synthesis in natural oil accumulations
e., hydrogenotrophic methanogenesis) is well recognized by petroleum/ remains unproven.
organic geochemists, the role of H2 in making oils is highly debatable. It has been hypothesized that H2 can originate from the thermal
Theories of abiotic (abiogenic) origin of oil rely on the availability of decay of organic matter in oil accumulations and then accumulate in
molecular H2 in the deep subsurface (e.g., Szatmari, 1989; Kenney et al., underlying water legs (Zinger, 1962). Some laboratory experiments
2002; Kutcherov and Krayushkin, 2010). However, the vast majority of indicate generation of H2 during thermal decay of kerogen (Bogomolov,
petroleum geoscientists and geochemists do not support theories of 1976; Li et al., 2015, 2017). There are examples of high concentration
abiotic origin of large natural accumulations of petroleum (Peters et al., (>10%) of free H2 in coal basins (see Table 25 in Zgonnik, 2020), pre­
2005; Glasby, 2006). sumably from coal metamorphism. Alternatively, this H2 could be
Based on our results (Figs. 9 and 15-16), H2 in conventional reser­ generated at the early stages of organic matter maturation and remained
voirs is mostly crustal-sourced. Meng et al. (2015) interpreted small trapped or adsorbed within the micropores of coals (Truche and
(<0.1%) concentrations of H2 in conventional oil and gas reservoirs in Bazarkina, 2019). Hydrogen can also originate from dehydrogenation of
the Bohai Bay Basin as mantle-derived, but our analysis in the contexts clay minerals that are abundant in many shale formations (Truche et al.,
of a large global dataset from various geological habitats suggests that 2018). Such shale formations, when enriched in organic matter and
H2 there is sourced from the crust. exposed to sufficiently high thermal stress (Pepper and Corvi, 1995),
There are some ideas and hypotheses about how molecular H2 can serve as petroleum source rocks. Carbonate minerals may preserve H2 in
contribute to formation and destruction of oils that originated from their crystalline structure (Levshounova, 1991) and release it during the
organic-rich petroleum source rocks (Zhu et al., 2017). For example, a maturation and expulsion of petroleum from carbonate source rocks.
prominent petroleum geologist of the first half of the 20th century, When some of the reservoir or source rocks contain evaporite/potash
Wallace E. Pratt, proposed a model in which early-expelled relatively deposits, radiolysis of water by the radioactive decay of potassium may
heavy (i.e., low API gravity) oils rich in “unsaturated” hydrocarbons contribute H2 to petroleum fluids (Smith et al., 2005; Gregory et al.,
turn into lighter oils more rich in “saturated” compounds via hydroge­ 2019). On the other hand, free H2 can result from the destruction of
nation in the reservoirs (Pratt, 1934). This model was not generally petroleum fluids, for example, it can be produced via radiolytic

19
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

dehydrogenation of crude oil molecules (Frolov et al., 1998; Larter et al., literature in some details: two in Kansas (Coveney Jr. et al., 1987) and
2019). Free H2 likely plays a role in oil biodegradation and biological one in Mali (Prinzhofer et al., 2018). The concept of “hydrogen system”
aromatization of oil (Röling et al., 2003). (Prinzhofer et al., 2018; Rezaee, 2021) can be formulated similarly to
the concept of “petroleum system” (Magoon and Dow, 1994), and as it
5.3. Exploration for geologic H2 was done for “helium system” (McDowell et al., 2017; Danabalan et al.,
2022). This study addresses the origins of molecular H2, and it appears
At present, geologic H2 is produced and used as a source of energy that sources of hydrocarbons and H2 in the subsurface are very different.
(electricity) only in the village of Bourakébougou in Mali (Hydroma, Further, mechanisms and efficiency of migration, accumulation and
2022). The discovery of a H2-rich gas accumulation there was seren­ preservation of molecular H2 (much smaller, mobile, and less soluble in
dipitous because the exploration well targeted water and not H2 (Briere water than hydrocarbon gases) may be significantly different from those
et al., 2017; Prinzhofer et al., 2018). Serendipity also plays a role in the for hydrocarbon gases.
discovery of petroleum accumulations (Milkov and Navidi, 2020), He-
rich accumulations and geothermal fields (Dobson, 2016). However, I 6. Conclusions
propose that specific exploration programs may be designed and carried
out in order to deliberately search for accumulations of natural H2-rich Geologic molecular H2 is most abundant in natural gases sampled
gases. from volcanic/magmatic vents at MORs and in serpentinites. Hydrogen
Areas where H2-rich gases seep at the surface may be attractive for isotope data may be used to distinguish broadly defined crustal (δ2H-H2
initial H2 exploration. This concept is the same as seep-guided petroleum values <-650‰), mantle and primordial (δ2H-H2 values >-650‰) H2 in
exploration in 1800s – early 1900s (Link, 1952). First commercial nat­ natural gases, and integration with geological data and isotope data on
ural gas production in the USA (Fredonia, NY) from the naturally frac­ other associated gases such as He and CH4 improves reliability of such
tured shale reservoir (Devonian Dunkirk Shale) in 1825 resulted from interpretations. Identification of more specific mechanisms of H2 gen­
the observations of nearby gas seeps (Lash and Lash, 2014). eration based on isotopic composition data remains problematic because
Onshore seeps of H2-rich gases occur in areas of serpentinites (Neal the empirical dataset of δ2H-H2 values is relatively small and there is
and Stranger, 1983; Etiope et al., 2011), which make these areas a much overlap of values resulting from different generation processes.
possible target for dedicated H2 exploration. Natural gases sampled from Further complications arise from variations in source water isotope
serpentinites contain not only H2, but also valuable CH4 (Table 1). It is composition and temperature-dependent equilibrium isotopic fraction­
possible to directly capture H2-rich gases from serpentinite seeps and ation and kinetic exchanges between H2 and water. It appears that H2
vents, as a H2 production mechanism. The concept is comparable to steel from serpentinization reactions has δ2H-H2 values around -715‰, and
pyramids/tents above seafloor seeps near Platform Holly in the Coal Oil H2 supplied by the deep Earth (perhaps directly from the deep mantle
Point field offshore California (Hornafius et al., 1999). H2-rich gases also and core) has δ2H-H2 values around -350‰. However, these δ2H-H2
can be recovered during degassing of water produced from wells drilled values may also result from the temperature-dependent isotopic frac­
into fractured serpentinites (Boulart et al., 2013; Suda et al., 2014). tionation between H2 and water. More field and experimental data is
Surface H2 surveys may help find areas with potential sub-surface H2 needed to constrain δ2H-H2 values typical for specific potentially
accumulations (Prinzhofer et al., 2019; Frery et al., 2021). Shallow important processes of H2 generation in the subsurface such as radiolysis
aquifers and maybe even free gas accumulations in fractured serpen­ and fracture-induced reduction of water and thermogenic liberation of
tinite reservoirs potentially can be identified with geophysical technol­ H2 during petroleum cracking and coal metamorphism. This review
ogies, for example, by searching for magnetic anomalies (magnetite is a highlights the possibility of purposeful exploration for geologic H2,
product of serpentinization). The technology of hydraulic fracturing that which is a valuable natural resource. Onshore serpentinites with known
revolutionized petroleum industry at the beginning of the 21st century surface and subsurface occurrences of H2-rich gases are viable settings
(King, 2012) may potentially be utilized in serpentinite reservoirs. for such exploration programs.
Projects of H2 production from serpentinites may be further enhanced by
CO2 sequestration in serpentinite aquifers through formation of car­ CRediT authorship contribution statement
bonate minerals (Power et al., 2013). Oman has well-documented sur­
face (Fritz et al., 1992) and subsurface (Boulart et al., 2013) occurrences Alexei V. Milkov: Conceptualization, Data curation, Formal anal­
of H2-rich gases in the Semail ophiolites and well-developed petroleum ysis, Investigation, Methodology, Writing – review & editing.
exploration and development infrastructure, which make this country a
potential candidate for favorable H2 exploration. Japan is another
country where H2 may be recovered from shallow aquifers within ser­ Declaration of Competing Interest
pentinized rocks. Degassing of H2-rich gas from water wells sourcing a
commercial spa may be already happening in the Hakuba Happo hot The author declares that he has no known competing financial in­
springs areas, as is illustrated by Suda et al. (2014) in their Fig. 2. terests or personal relationships that could have appeared to influence
Extraction from waters producing geothermal energy may be the work reported in this paper.
another viable option for economic production of geologic H2. Subsur­
face volcanic/magmatic hydrothermal systems produce gases with sig­ Acknowledgments
nificant H2 (average ~7%, on water-free basis). Examples of systems
where H2 can be recovered during production of geothermal energy The author would like to thank Antonio Martín-Monge, Geoffrey
include such H2-rich geothermal fields as Salton Sea (Welhan, 1981) in Ellis, Chris Boreham and an anonymous reviewer for providing useful
the USA, Cerro Prieto in Mexico (Welhan, 1981), and Namafjall (Sano comments on the manuscript.
et al., 1985), Nesjavellir (Marty et al., 1991) and Hellisheidi (Combau­
don et al., 2022) in Iceland. Subsurface sedimentary hydrothermal sys­
References
tems produce gases with relatively less H2 (average ~4%, on water-free
basis), and may be less attractive targets for H2 exploration. Abrajano, T.A., Sturchio, N.C., Bohlke, J.K., Lyon, G.L., Poreda, R.J., Stevens, C.M., 1988.
Petroleum companies may be more comfortable with the idea of Methane-hydrogen gas seeps, Zambales Ophiolite, Philippines: deep or shallow
exploring for H2-rich gases in conventional reservoirs. There are known origin? Chem. Geol. 71, 211–222.
Adam, N., Perner, M., 2018. Microbially mediated hydrogen cycling in deep-sea
accumulations of H2-rich gas in conventional reservoirs (as reviewed by hydrothermal vents. Front. Microbiol. 9, 2873. https://doi.org/10.3389/
Zgonnik, 2020), but only three such gas pools have been described in the fmicb.2018.02873.

20
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Adams, M.W.W., Stiefel, E.I., 1988. Biological hydrogen production: not so elementary. Bouquet, A., Glein, C.R., Wyrick, D., Waite, J.H., 2017. Alternative energy: production of
Science 282, 1842–1843. https://doi.org/10.1126/science.282.5395.1842. H2 by radiolysis of water in the rocky cores of icy bodies. The Astrophys. J. Lett. 840,
Albers, E., Bach, W., Pérez-Gussinyé, M., McCammon, C., Frederichs, T., 2021. L8. https://doi.org/10.3847/2041-8213/aa6d56.
Serpentinization-driven H2 production from continental break-up to Mid-Ocean Boyd, A.W., Willis, C., Miller, O.A., 1973. The radiolysis of water vapor at very high dose
Ridge spreading: unexpected high rates at the West Iberia Margin. Front. Earth Sci. rates. I. Hydrogen yields from H2O, H2O-HCl, and H2O-HBr mixtures. Can. J. Chem.
9, 673063 https://doi.org/10.3389/feart.2021.673063. 51, 4048–4055. https://doi.org/10.1139/v73-603.
Anderson, R.B., 1984. The Fischer-Tropsch synthesis. Academic Press, New York. Briere, D., Jerzykiewicz, T., Sliwinski, W., 2017. On generating a geological model for
Angino, E.E., Coveney, R.M.J., Goebel, E.D., Zeller, E.J., Dreschhoff, G.A.M., 1984. hydrogen gas in the southern Taoudeni Megabasin (Bourakebougou area, Mali). In:
Hydrogen and nitrogen – origin, distribution, and abundance, a followup. Oil Gas J. Search and Discovery Article #42041. Available online at. https://www.searcha
82, 142–146. nddiscovery.com/pdfz/documents/2017/42041jerzykiewicz/ndx_jerzykiewicz.pdf.
Apps, J.A., van de Kamp, P.C., 1993. Energy gases of abiogenic origin in the Earth’s crust. html.
In: Howell, D.G. (Ed.), The Future of Energy Gases. United States Geological Survey Brophy, J.G., Schimmelmann, A., 2017. Hydrogen. In: White, W.M. (Ed.), Encyclopedia
Professional Paper 1570, Washington, pp. 81–130. of Geochemistry. Springer. https://doi.org/10.1007/978-3-319-39193-9_325-1.
Arai, T., Okusawa, T., Tsukahara, H., 2001. Behavior of gases in the Nojima Fault Zone Online ISBN 978-3-319-39193-9.
revealed from the chemical composition and carbon isotope ratio of gases extracted Byrne, D.J., Barry, P.H., Lawson, M., Ballentine, C.J., 2018. Noble gases in conventional
from DPRI 1800 m drill core. Island Arc 10, 430–438. https://doi.org/10.1111/ and unconventional petroleum systems. Geol. Soc. Lond. Spec. Publ. 468 (1),
j.1440-1738.2001.00341.x. 127–149.
Arnason, B., 1977. The hydrogen-water isotope thermometer applied to geothermal areas Byrne, D.J., Broadley, M.W., Halldórsson, S.A., Ranta, E., Ricci, A., Tyne, R.L.,
in Iceland. Geotherm. 5, 75–80. Stefánsson, A., Ballentine, C.J., Barry, P.H., 2021. The use of noble gas isotopes to
Arnason, B., Sigurgeirsson, T., 1968. Deuterium content of water vapour and hydrogen in trace subsurface boiling temperatures in Icelandic geothermal systems. Earth Planet.
volcanic gas at Surtsey, Iceland. Geochim. Cosmochim. Acta 32, 807–813. Sci. Lett. 560, 116805 https://doi.org/10.1016/j.epsl.2021.116805.
Arrouvel, C., Prinzhofer, A., 2021. Genesis of natural hydrogen: new insights from Caby, R., 2014. Nature and evolution of Neoproterozoic ocean-continent transition:
thermodynamic simulations. Int. J. Hydrog. Energy 46, 18780–18794. https://doi. evidence from the passive margin of the West African craton in NE Mali. J. Afr. Earth
org/10.1016/j.ijhydene.2021.03.057. Sci. 91, 1–11. https://doi.org/10.1016/j.jafrearsci.2013.11.004.
Bannerjee, M., Mendhe, V.A., Kamble, A.D., Varma, A.K., Singh, B.D., Kumar, S., 2021. Campbell, J.H., Koskinas, G.J., Gallegos, G., Gregg, M., 1980. Gas evolution during oil
Facets of coalbed methane reservoir in East Bokaro Basin, India. J. Petr. Sci. Eng. shale pyrolysis. 1. Nonisothermal rate measurements. Fuel 59, 718–725.
109255 https://doi.org/10.1016/j.petrol.2021.109255. Chamberlin, R.T., 1909. The gases in rocks. The J. Geol. 17, 534–568.
Bao, Y., Wei, C., Wang, C., Wang, G., Li, Q., 2014. Geochemical characteristics and Charlou, J.L., Donval, J.P., Douville, E., Jean-Baptiste, P., Radford-Knoery, J.,
generation process of mixed biogenic and thermogenic coalbed methane in Luling Fouquet, Y., Dapoigny, A., Stievenard, M., 2000. Compared geochemical signatures
Coalfield, China. Energy Fuel 28, 4392–4401. and the evolution of Menez Gwen (37◦ 50′ N) and Lucky Strike (37◦ 17′ N)
Barbier, S., Huang, F., Andreani, M., Tao, R., Hao, J., Eleish, A., Prabhu, A., Minhas, O., hydrothermal fluids, south of the Azores Triple Junction on the Mid-Atlantic Ridge.
Fontaine, K., Fox, P., Daniel, I., 2020. A review of H2, CH4, and hydrocarbon Chem. Geol. 171, 49–75. https://doi.org/10.1016/S0009-2541(00)00244-8.
formation in experimental serpentinization using network analysis. Front. Earth Sci. Charlou, J.L., Donval, J.P., Fouquet, Y., Jean-Baptiste, P., Holm, N., 2002. Geochemistry
8, 209. https://doi.org/10.3389/feart.2020.00209. of high H2 and CH4 vent fluids issuing from ultramafic rocks at the Rainbow
Barnes, I., O’Neil, J.R., Trescases, J.J., 1978. Present day serpentinization in New hydrothermal field (36 14 N, MAR). Chem. Geol. 191, 345–359.
Caledonia, Oman and Yugoslavia. Geochim. Cosmochim. Acta 42, 144–145. Chatellier, J.Y., Ferworn, K., Lazreg Larsen, N., Ko, S., Flek, P., Molgat, M., Anderson, I.,
Behar, F., Vandenbroucke, M., Tang, Y., Marquis, F., Espitalié, J., 1997. Thermal 2013. Overpressure in shale gas: when geochemistry and reservoir engineering data
cracking of kerogen in open and closed systems: determination of kinetic parameters meet and agree. In: Chatellier, J., Jarvie, D. (Eds.), Critical Assessment of Shale
and stoichiometric coefficients for oil and gas generation. Org. Geochem. 26, Resource Plays, AAPG Memoir, 103, pp. 45–69.
321–339. Chen, Z., Chen, L., Wang, G., Zou, C., Jiang, S., Si, Z., Gao, W., 2020. Applying isotopic
Bergfeld, D., Lowenstern, J.B., Hunt, A.G., Shanks III, W.C.P., Evans, W.C., 2014. Gas and geochemical proxy for gas content prediction of Longmaxi shale in the Sichuan
isotope chemistry of thermal features in Yellowstone National Park, Wyoming (ver. Basin, China. Mar. Pet. Geol. 116, 104329 https://doi.org/10.1016/j.
1.1, September 2014). In: US Geol. Surv. Scient. Investig. Rep. 2011-5012, 28 p. and marpetgeo.2020.104329.
data files. https://doi.org/10.3133/sir20115012. Choi, J., Kim, J.-H., Kim, G.-Y., Chang, S.-W., 2016. Data report: gas isotope and
Bernard, B., Brooks, J.M., Sackett, W.M., 1977. A geochemical model for characterization conversion ratio of headspace and void gas, IODP Expedition 344. In: Harris, R.N.,
of hydrocarbon gas sources in marine sediments. In: 9th Annual Offshore Technology Sakaguchi, A., Petronotis, K., the Expedition 344 Scientists (Eds.), Proceedings of the
Conference, 435-438. OTC 2934. Integrated Ocean Drilling Program, 344: College Station, TX (Integrated Ocean
Bernard, B.B., Brooks, J.M., Orange, D.L., Decker, J., 2013. Interstitial light hydrocarbon Drilling Program). https://doi.org/10.2204/iodp.proc.344.203.2016.
gases in jumbo piston cores offshore Indonesia: thermogenic or biogenic?. In: Paper Combaudon, V., Moretti, I., Kleine, B.I., Stefánsson, A., 2022. Hydrogen emissions from
Presented at the Offshore Technology Conference, Houston, Texas, USA, May 2013. hydrothermal fields in Iceland and comparison with the Mid-Atlantic Ridge. Intern.
OTC 24228 https://doi.org/10.4043/24228-MS. J. Hydrog. Ener. 47, 10217–10227. https://doi.org/10.1016/j.
Bjornstad, B.N., McKinley, J.P., Stevens, T.O., Rawson, S.A., Fredrickson, J.K., Long, P.E., ijhydene.2022.01.101.
1994. Generation of hydrogen gas as a result of drilling within the saturated zone. Conrad, R., 1995. Soil microbial processes involved in production and consumption of
Groundw. Monit. Remed. 14, 140–147. https://doi.org/10.1111/j.1745-6592.1994. atmospheric trace gases. Adv. Microb. Ecol. 14, 207–250.
tb00492.x. Conrad, R., 1996. Soil microorganisms as controllers of atmospheric trace gases (H2, CO,
Blair, C.C., D’Hondt, S., Spivack, A.J., Kingsley, R.H., 2007. Radiolytic hydrogen and CH4, OCS, N2O, and NO). Microbiol. Rev. 60, 609–640.
microbial respiration in subsurface sediments. Astrobiology 7, 951–970. Coveney Jr., R.M., Goebel, E.D., Zeller, E.J., Dreschhoff, G.A.M., Angino, E.E., 1987.
Bogomolov, A., 1976. Organic matter of sedimentary rocks as a source of hydrogen in Serpentinization and the origin of hydrogen gas in Kansas. AAPG Bull. 71, 39–48.
natural gases. In: Studies of Organic Matter in Modern and Ancient Sediments. https://doi.org/10.1306/94886D3F-1704-11D7-8645000102C1865D.
Nauka, Moscow, pp. 79–80. Craig, H., Lupton, J.E., 1976. Primordial neon, helium, and hydrogen in oceanic basalts.
Bohdanowicz, C., 1934. Natural gas occurrences in Russia (USSR). AAPG Bull. 18, Earth Planet. Sci. Lett. 31, 369–385.
750–760. Crespo-Medina, M., Twing, K., Sánchez, R., Brazelton, W.J., McCollom, T.M., Screnk, M.
Bokova, E.N., 1953. Formation of methane during microbial degradation of oil. Polev. O., 2017. Methane dynamics in a tropical serpentinizing environment: the Santa
Promysl. Geochim. 2, 25–27 (in Russian). Elena Ophiolite, Costa-Rica. Front. Microbiol. 9, 916. https://doi.org/10.3389/
Boone, D.R., Johnson, R.L., Liu, Y., 1989. Diffusion of the interspecies electron carriers fmicb.2017.00916.
H2 and formate in methanogenic ecosystems and its implications in the measurement Daae, F.L., Økland, I., Dahle, H., Jørgensen, S.L., Thorseth, I.H., Pedersen, R.B., 2013.
of Km for H2 or formate uptake. Appl. Environ. Microbiol. 55, 1735–1741. https:// Microbial life associated with low-temperature alteration of ultramafic rocks in the
doi.org/10.1128/AEM.55.7.1735-1741.1989. Leka ophiolite complex. Geobiology 11, 318–339. https://doi.org/10.1111/
Boreham, C.J., Davies, J.B., 2020. Carbon and hydrogen isotopes of the wet gases gbi.12035.
produced by gamma-ray-induced polymerisation of methane: insights into D’Alessandro, W., Yüce, G., Italiano, F., Bellomo, S., Gülbay, A.H., Yasin, D.U.,
radiogenic mechanism and natural gas formation. Radiat. Phys. Chem. 168, 108546 Gagliano, A.L., 2018. Large compositional differences in the gases released from the
https://doi.org/10.1016/j.radphyschem.2019.108546. Kizildag ophiolitic body (Turkey): evidences of prevailingly abiogenic origin. Mar.
Boreham, C.J., Edwards, D.S., Czado, K., Rollet, N., Wang, L., van der Wielen, S., Pet. Geol. 89, 174–184. https://doi.org/10.1016/j.marpetgeo.2016.12.017.
Champion, D., Blewett, R., Feitz, A., Henson, P.A., 2021a. Hydrogen in Australian Danabalan, D., Gluyas, J.G., Macpherson, C.G., Abraham-James, T.H., Bluett, J.J.,
natural gas: occurrences, sources and resources. The APPEA J. 61, 163–191. https:// Barry, P.H., Ballentine, C.J., 2022. The principles of helium exploration. Pet. Geosci.
doi.org/10.1071/AJ20044. 28 https://doi.org/10.1144/petgeo2021-029 petgeo2021–029.
Boreham, C.J., Sohn, J.H., Cox, N., Williams, J., Hong, Z., Kendrick, M.A., 2021b. Darrah, T.H., Vengosh, A., Jackson, R.B., Warner, N.R., Poreda, R.J., 2014. Noble gases
Hydrogen and hydrocarbons associated with the Neoarchean Frog’s Leg Gold Camp, identify the mechanisms of fugitive gas contamination in drinking-water wells
Yilgarn Craton, Western Australia. Chem. Geol. 120098 https://doi.org/10.1016/j. overlying the Marcellus and Barnett shales. Proc. Natl. Acad. Sci. 111,
chemgeo.2021.120098. 14,076–14,081. https://doi.org/10.1073/pnas.1322107111.
Boschetti, T., Etiope, G., Toscani, L., 2013. Abiotic methane in the hyperalkaline springs Daskalopoulou, K., Calabrese, S., Grassa, F., Kyriakopoulos, K., Parello, F., Tassi, F.,
of Genova, Italy. Proc. Earth Planet. Sci. 7, 248–251. D’Alessandro, W., 2018. Origin of methane and light hydrocarbons in natural fluid
Boulart, C., Chavagnac, V., Monnin, C., Delacour, A., Ceuleneer, G., Hoareau, G., 2013. emissions: a key study from Greece. Chem. Geol. 479, 286–301.
Differences in gas venting from ultramafic-hosted warm springs: The example of Dauphas, N., Robert, F., Marty, B., 2000. The late asteroidal and cometary bombardment
Oman and Voltri ophiolites. Ofioliti 38, 142–156. https://doi.org/10.4454/ofioliti. of Earth as recorded in water deuterium to protium ratio. Icarus 148, 508–512.
v38i2.423.

21
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Deloule, E., Albarede, F., Sheppard, S.M.F., 1991. Hydrogen isotope heterogeneities in Freund, F., Dickinson, J.T., Cash, M., 2002. Hydrogen in rocks: an energy source for deep
the mantle from ion probe analysis of amphiboles from ultramafic rocks. Earth microbial communities. Astrobiology 2, 83–92. https://doi.org/10.1089/
Planet. Sci. Lett. 105, 543–553. 153110702753621367.
Deshaee, A., Shakeria, A., Taran, Y., Mehrabi, B., Farhadian, M., Zelenski, M., Fridman, A., 1970. Natural gases of ore fields. Nedra, Moscow (in Russian).
Chaplygin, I., Tassi, F., 2020. Geochemistry of Bazman thermal springs, southeast Fritz, P., Clark, I.D., Fontes, J.C., Whiticar, M.J., Faber, E., 1992. Deuterium and 13C
Iran. J. Volcanol. Geotherm. Res. 390, 106676 https://doi.org/10.1016/j. evidence for low temperature production of hydrogen and methane in a highly
jvolgeores.2019.106676. alkaline groundwater environment in Oman. In: Kharaka, Y., Maest, A.S. (Eds.),
Deville, E., Prinzhofer, A., 2016. The origin of N2-H2-CH4-rich natural gas seepages in Proceedings of the 7th International Symposium on Water-Rock Interaction.
ophiolitic context: a major and noble gases study of fluid seepages in New Caledonia. Rotterdam Balkama, pp. 792–796.
Chem. Geol. 440, 139–147. https://doi.org/10.1016/j.chemgeo.2016.06.011. Frolov, E.B., Smirnov, M.B., Melikhov, V.A., Vanyukova, N.A., 1998. Olefins of
Dick, H., Lin, J., Schouten, H., 2003. An ultraslow-spreading class of ocean ridge. Nature radiogenic origin in crude oils. Org. Geochem. 29, 409–420.
426, 405–412. https://doi.org/10.1038/nature02128. Fryer, P., 2012. Serpentinite mud volcanism: observations, processes, and implications.
Dobson, P.F., 2016. A review of exploration methods for discovering hidden geothermal Annu. Rev. Mar. Sci. 4, 345–373.
systems. GRC Transact. 40, 695–706. Galimov, E.M., Petersil’ye, I.A., 1967. Isotopic composition of carbon from hydrocarbon
Drobner, E., Huber, H., Wächterschäuser, G., Rose, D., Stetter, K.O., 1990. Pyrite gases and CO2 held in alkalic igneous rocks of the Khibiny, Lovozero and Ilímaussak
formation linked with hydrogen evolution under anaerobic conditions. Nature 346, plutons. Dokl. Akad. Nauk SSSR 176, 200–203.
742–744. https://doi.org/10.1038/346742a0. Galimov, E.M., Prokhorov, V.S., Fedoseyev, D.V., Varnin, V.P., 1973. Heterogenous
Dubessy, J., Ramboz, C., 1986. The history of organic nitrogen from early diagenesis to carbon isotope effects in synthesis of diamond and graphite from gas. Geokhimiya 3,
amphibolite facies: mineralogical, chemical, mechanical and isotopic consequences. 416–424.
In: Vth International Symposium on Water-Rock Interaction. Reykjavik, Iceland, Gasperini, L., Lazar, M., Mazzini, A., Lupi, M., Haddad, A., Hensen, C., Schmidt, M.,
pp. 171–174. Extended abstracts. Caracausi, A., Ligi, M., Polonia, A., 2020. Neotectonics of the Sea of Galilee
Dubessy, J., Pagel, M., Beny, J.M., Christensen, H., Hickel, B., Kosztolanyi, C., Poty, B., (northeast Israel): implication for geodynamics and seismicity along the Dead Sea
1988. Radiolysis evidenced by H2-O2 and H2-bearing fluid inclusions in three Fault system. Sci. Rep. 10, 11932. https://doi.org/10.1038/s41598-020-67930-6.
uranium deposits. Geochem. Cosmochim. Acta 52, 1155–1167. Gaucher, E.C., 2020. New perspectives in the industrial exploration of native hydrogen.
Dubessy, J., Poty, B., Ramboz, C., 1989. Advances in C-O-H-N-S fluid geochemistry based Elements 16, 8–9.
on micro-Raman spectrometric analysis of fluid inclusions. Eur. J. Mineral. 1, Gest, H., 1954. Oxidation and evolution of molecular hydrogen by microorganisms.
517–534. https://doi.org/10.1127/ejm/1/4/0517. Bacteriol. Rev. 18, 43–73.
Dzaugis, M.E., Spivack, A.J., Dunlea, A.G., Murray, R.W., D’Hondt, S., 2016. Radiolytic Geymond, U., Ramanaidou, E., Lévy, D., Ouaya, A., Moretti, I., 2022. Can weathering of
hydrogen production in the subseafloor basaltic aquifer. Front. Microbiol. 7, 76. banded iron formations generate natural hydrogen? Evidence from Australia, Brazil
https://doi.org/10.3389/fmicb.2016.00076. and South Africa. Minerals 12, 163. https://doi.org/10.3390/min12020163.
Ehhalt, D.H., Rohrer, F., 2009. The tropospheric cycle of H2: a critical review. Tellus Ser. Gilat, A.L., Vol, A., 2005. Primordial hydrogen-helium degassing, an overlooked major
B Chem. Phys. Meteorol. 61, 500–535. https://doi.org/10.1111/j.1600- energy source for internal terrestrial processes. HAIT J. Sci. Eng. B 2, 125–167.
0889.2009.00416.x. Gilat, A.L., Vol, A., 2012. Degassing of primordial hydrogen and helium as the major
Ekzercev, V.A., 1960. Formation of methane by microorganisms in oil fields. Geokhimiya energy source for internal terrestrial processes. Geosci. Front. 3, 911–921.
1, 362–370 (in Russian). Ginsburg-Karagitscheva, T.L., 1933. Microflora of oil waters and oil-bearing formations
Erzinger, J., Wiersberg, T., Zimmer, M., 2006. Real-time mud gas logging and sampling and biochemical processes caused by it. AAPG Bull. 17, 52–65.
during drilling. Geofluids 6, 225–233. https://doi.org/10.1111/j.1468- Glasby, G.P., 2006. Abiogenic origin of hydrocarbons: an historical overview. Resour.
8123.2006.00152.x. Geol. 56, 83–96.
Etiope, G., Schoell, M., 2014. Abiotic gas: atypical but not rare. Elements 10, 291–296. Glein, C.R., Zolotov, M.Yu., 2020. Hydrogen, hydrocarbons, and habitability across the
Etiope, G., Sherwood Lollar, B., 2013. Abiotic methane on Earth. Rev. Geophys. 51, solar system. Elements 16, 47–52.
276–299, 2012RG000428. 8755-1209/13/10.1002/rog.20011. Glindermann, D., Edwards, M., Morgenstern, P., 2005. Phosphine from rocks:
Etiope, G., Whiticar, M.J., 2019. Abiotic methane in continental ultramafic rock systems: mechanically driven phosphate reduction? Environ. Sci. Technol. 39, 7671–7675.
towards a genetic model. Appl. Geochem. 102, 139–152. https://doi.org/10.1016/j. https://doi.org/10.1021/es050682w.
apgeochem.2019.01.012. Goebel, E.D., Coveney Jr., R.M., Angino, E.E., Zeller, E.J., Dreschhoff, G.A.M., 1984.
Etiope, G., Feyzullayev, A., Milkov, A.V., Waseda, A., Mizobe, K., Sun, C.H., 2009. Geology, composition, isotopes of naturally occurring H2/N2 rich gas from wells near
Evidence of subsurface anaerobic biodegradation of hydrocarbons and potential Junction City. Kansas. Oil Gas J. 82, 215–222.
secondary methanogenesis in terrestrial mud volcanoes. Mar. Pet. Geol. 26, Goetz, A.M., 2021. Regional Groundwater Conditions in Northeast BC: Results from a
1692–1703. Monitoring Well Network in an Area of Historical and Ongoing Unconventional
Etiope, G., Schoell, M., Hosgörmez, H., 2011. Abiotic methane flux from the Chimaera Natural Gas Development. MSc Thesis. The University of British Colombia, p. 145.
seep and Tekirova ophiolites (Turkey): understanding gas exhalation from low Gordienko, V., 2019. On hydrogen degassing in the areas of recent activation of Ukraine.
temperature serpentinization and implications for Mars. Earth Planet. Sci. Lett. 310, Geofiz. Zhurn. 41, 115–127 (in Russian). 10.24028/gzh.0203-3100.v41i5.2019.1
96–104. https://doi.org/10.1016/j.epsl.2011.08.001. 83617.
Etiope, G., Vance, S., Christensen, L.E., Marques, J.M., Ribeiro da Costa, I., 2013. Gregory, S.P., Barnett, M.J., Field, L.P., Milodowski, A.E., 2019. Subsurface microbial
Methane in serpentinized ultramafic rocks in mainland Portugal. Mar. Pet. Geol. 45, hydrogen cycling: natural occurrence and implications for industry. Microorganisms
12–16. https://doi.org/10.1016/j.marpetgeo.2013.04.009. 7, 53. https://doi.org/10.3390/microorganisms7020053.
Etiope, G., Vadillo, I., Whiticar, M.J., Marques, J.M., Carreira, P.M., Tiago, I., Gruca-Rokosz, R., Koszelnik, P., 2018. Production pathways for CH4 and CO2 in
Benavente, J., Jiménez, P., Urresti, B., 2016. Abiotic methane seepage in the Ronda sediments of two freshwater ecosystems in south-eastern Poland. PLoS One 13,
peridotite massif, southern Spain. Appl. Geochem. 66, 101–113. https://doi.org/ e0199755. https://doi.org/10.1371/journal.pone.0199755.
10.1016/j.apgeochem.2015.12.001. Guélard, J., Beaumont, V., Rouchon, V., Guyot, F., Pillot, D., Jezequel, D., Ader, M.,
Etiope, G., Samardzic, N., Grassa, F., Hrvatovic, H., Miosic, N., Skopljak, F., 2017. Newell, K.D., Deville, E., 2017. Natural H2 in Kansas: deep or shallow origin?
Methane and hydrogen in hyperalkaline groundwaters of the serpentinized Dinaride Geochem. Geophys. Geosyst. 18, 1841–1865. https://doi.org/10.1002/
ophiolite belt, Bosnia and Herzegovina. Appl. Geochem. 84, 286–296. 2016GC006544.
Evans, B.W., Hattori, K., Baronnet, A., 2013. Serpentinite: what, why, where? Elements Guillot, S., Hattori, K., 2013. Serpentinites: essential roles in geodynamics, arc
9, 99–106. https://doi.org/10.2113/gselements.9.2.99. volcanism, sustainable development, and the origin of life. Elements 9, 95–98.
Farias, C.B.B., Barreiros, R.C.S., da Silva, M.F., Casazza, A.A., Converti, A., Sarubbo, L.A., Hachikubo, A., Minami, H., Yamashita, S., Khabuev, A., Krylov, A., Kalmychkov, G.,
2022. Use of hydrogen as fuel: a trend of the 21st century. Energies 15, 311. https:// Poort, J., De Batist, M., Chenskiy, A., Manakov, A., Khlystov, O., 2020.
doi.org/10.3390/en15010311. Characteristics of hydrate-bound gas retrieved at the Kedr mud volcano (southern
Feldbusch, E., Wiersberg, T., Zimmer, M., Regenspurg, S., 2018. Origin of gases from the Lake Baikal). Sci. Rep. 10, 14747.
geothermal reservoir Groß Schӧnebeck (North German Basin). Geothermics 71, Hahn-Weinheimer, P., Hirner, A., 1981. Isotopic evidence for the origin of graphite.
357–368. https://doi.org/10.1016/j.geothermics.2017.09.007. Geochem. J. 15, 9–15.
Feng, Z., Huang, S., Wu, W., Xie, C., Peng, W., Cai, Y., 2017. Longmaxi shale gas Halas, P., Dupuy, A., Franceschi, M., Boardmann, V., Fleury, J.-M., 2021. Hydrogen gas
geochemistry in Changning and Fuling gas fields, the Sichuan Basin. Energy Explor. in circular depressions in South Gironde, France: flux, stock, or artefact? Appl.
Exploit. 35, 259–278. Geochem. 127, 104928 https://doi.org/10.1016/j.apgeochem.2021.104928.
Fiebig, J., Stefánsson, A., Ricci, A., Tassi, F., Viveiros, F., Silva, C., Lopez, T.M., Hallenbeck, P.C., Benemann, J.R., 2002. Biological hydrogen production; fundamentals
Schreiber, C., Hofmann, S., Mountain, B.W., 2019. Abiogenesis not required to and limiting processes. Int. J. Hydrog. Energy 27, 1185–1193. https://doi.org/
explain the origin of volcanic-hydrothermal hydrocarbons. Geochem. Persp. Let. 11, 10.1016/S0360-3199(02)00131-3.
23–27. Hanley, E.S., Deane, J.P., Gallachóir, B.P.Ó., 2018. The role of hydrogen in low carbon
Fischer, T.P., Chiodini, G., 2015. Volcanic, magmatic and hydrothermal gases. In: energy futures - a review of existing perspectives. Renew. Sust. Energ. Rev. 82,
Sigurdsson, H. (Ed.), The Encyclopedia of Volcanoes (Second Edition). Academic 3027–3045.
Press, pp. 779–797. https://doi.org/10.1016/B978-0-12-385938-9.00045-6. Hao, Y., Pang, Z., Tian, J., Wang, Y., Li, Z., Li, L., 2020. Origin and evolution of
Fischer, F., Tropsch, H., 1926. The synthesis of petroleum at atmospheric pressures from hydrogen-rich gas discharges from a hot spring in the eastern coastal area of China.
gasification products of coal. Brennstoff-Chemie 7, 97–104. Chem. Geol. 538, 119477 https://doi.org/10.1016/j.chemgeo.2020.119477.
Frery, E., Langhi, L., Maison, M., Moretti, I., 2021. Natural hydrogen seeps identified in Hawkes, H.E., 1972. Free hydrogen in genesis of petroleum. AAPG Bull. 56, 2268–2277.
the North Perth Basin, Western Australia. Int. J. Hydrog. Energy 46, 31158–31173. He, Y., Sun, S., Kim, D.Y., Jang, B.G., Li, H., Mao, H., 2022. Superionic iron alloys and
https://doi.org/10.1016/j.ijhydene.2021.07.023. their seismic velocities in Earth’s inner core. Nature 602, 258–262. https://doi.org/
10.1038/s41586-021-04361-x.

22
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Headlee, A.J.W., 1962. Hydrogen sulfide, free hydrogen are vital exploration clues. Kita, I., Matsuo, S., Wakita, H., Nakamura, Y., 1980. D/H ratios of H2 in soil gases as an
World Oil, November, pp. 78–83. indicator of fault movements. Geochem. J. 14, 317–320.
Hendrix, M.S., 2009. Continental sediments. In: Gornitz, V. (Ed.), Encyclopedia of Kita, I., Matsuo, S., Wakita, H., 1982. H2 generation by reaction between H2O and
Paleoclimatology and Ancient Environments. Encyclopedia of Earth Sciences Series. crushed rock: an experimental study on H2 degassing from the active fault zone.
Springer, Dordrecht. https://doi.org/10.1007/978-1-4020-4411-3_47. J. Geophys. Res. 87, 10789–10795. https://doi.org/10.1029/JB087iB13p10789.
Hirose, T., Kawagucci, S., Suzuki, K., 2011. Mechanoradical H2 generation during Kiyosu, Y., 1983. Hydrogen isotopic compositions of hydrogen and methane from some
simulated faulting: implications for an earthquake-driven subsurface biosphere. volcanic areas in northeastern Japan. Earth Planet. Sci. Lett. 62, 41–52. https://doi.
Geophys. Res. Lett. 39, L23304. https://doi.org/10.1029/2012GL054539. org/10.1016/0012-821X(83)90069-9.
Hirose, T., Kawagucci, S., Suzuki, K., 2012. Correction to Mechanoradical H2 generation Klein, F., Bach, W., McCollom, T.M., 2013. Compositional controls on hydrogen
during simulated faulting: implications for an earthquake-driven subsurface generation during serpentinization of ultramafic rocks. Lithos 178, 55–69.
biosphere. Geophys. Res. Lett. 39, L054539. https://doi.org/10.1029/ Klein, F., Tarnas, J.D., Bach, W., 2020. Abiotic sources of molecular hydrogen on Earth.
2012GL054539. Elements 16, 19–24. https://doi.org/10.2138/gselements.16.1.19.
Hoeman, E.C., 1944. The Combustion of Phosphorus. PhD Thesis. Missouri University of Komabayashi, T., 2021. Hydrogen dances in the deep mantle. Nature Geosc. 14,
Science and Technology, USA. Available at. https://scholarsmine.mst.edu/profe 112–117.
ssional_theses/284/ (accessed 15 December 2021). Konn, C., Donval, J.P., Guyader, V., Germain, Y., Alix, A.-S., Roussel, E., Rouxel, O.,
Hofstra, A.H., Cline, J.S., 2000. Characteristics and models for Carlin-type gold deposits. 2022. Extending the dataset of fluid geochemistry of the Menez Gwen, Lucky Strike,
Rev. Econ. Geol. 13, 163–220. Rainbow, TAG and Snake Pit hydrothermal vent fields: Investigation of temporal
Holland, H.D., 2002. Volcanic gases, black smokers, and the great oxidation event. stability and organic contribution. Deep Sea Res. Part I: Ocean. Res. Papers 179,
Geochim. Cosmochim. Acta 66, 3811–3826. https://doi.org/10.1016/S0016-7037 103630. https://doi.org/10.1016/j.dsr.2021.103630.
(02)00950-X. Kraychuk, M.S., 1976. About hydrogen gas in the Gagarin prospect. Oil-Gas Geol.
Holm, N.G., Oze, C., Mousis, O., Waite, J.H., Guilbert-Lepoutreet, A., 2015. Geophys. 19, 1–10.
Serpentinization and the formation of H2 and CH4 on celestial bodies (planets, Krichevsky, M.I., Friedman, I., Newell, M.F., Sisler, F.D., 1961. Deuterium fractionation
moons, comets). Astrobiology 15, 587–600. during molecular hydrogen formation in a marine pseudomonad. J. Biol. Chem. 236,
Horibe, Y., Craig, H., 1995. D/H fractionation in the system methane-hydrogen-water. 2520–2525. https://doi.org/10.1016/S0021-9258(18)64032-3.
Geochim. Cosmochim. Acta 59, 5209–5217. Kutcherov, V.G., Krayushkin, V.A., 2010. Deep-seated abiogenic origin of petroleum:
Hornafius, J.S., Quigley, D., Luyendyk, B.P., 1999. The world’s most spectacular marine from geological assessment to physical theory. Rev. Geophys. 48, RG1001.
hydrocarbon seeps (Coal Oil Point, Santa Barbara Channel, California): Kyser, T.K., O’Neil, J.R., 1984. Hydrogen isotope systematics of submarine basalts.
quantification of emissions. J. Geophys. Res. 104, 20703–20711. https://doi.org/ Geochim. Cosmochim. Acta 48, 2123–2133.
10.1029/1999JC900148. Lang, S.Q., Brazelton, W.J., 2020. Habitability of the marine serpentinites subsurface: a
Horsfield, B., Mahlstedt, N., Weniger, P., Misch, D., Vranjes-Wessely, S., Han, S., case study of the Lost City hydrothermal field. Phil. Trans. R. Soc. A 378, 20180429.
Wang, C., 2022. Molecular hydrogen from organic sources in the deep Songliao https://doi.org/10.1098/rsta.2018.042.
Basin, P.R China. Int. J. Hydrog. Energy 47, 16750–16774. https://doi.org/10.1016/ Larin, V.N., 1993. Hydridic Earth: the New Geology of our Primordially Hydrogen-Rich
j.ijhydene.2022.02.208. Planet. Polar Publishing, Alberta, Canada.
Huang, F., Barbier, S., Tao, R., Hao, J., del Real, P.G., Peuble, S., Merdith, A., Larin, V.N., 1995. Our Earth (Origin, Composition, Structure and Development of
Leichnig, V., Perrillat, J.-P., Fontaine, K., Fox, P., Andreani, M., Daniel, I., 2019. Initially Hydridic Earth). Agar, Moscow (in Russian).
Dataset for H2, CH4 and organic compounds formation during experimental Larin, N.V., Zgonnik, V., Rodina, S., Deville, E., Prinzhofer, A., Larin, V.N., 2015. Natural
serpentinization. Geosc. Data J. 8, 90–100. https://doi.org/10.1002/gdj3.105. molecular hydrogen seepage associated with surficial, rounded depressions on the
Hunt, J.M., 1996. Petroleum Geochemistry and Geology. W.H. Freeman and Co., New European craton in Russia. Nat. Resour. Res. 24, 369–383. https://doi.org/10.1007/
York. s11053-014-9257-5.
Hydroma, 2022. https://hydroma.ca/en/natural-hydrogen/. Larter, S., Silva, R.C., Marcano, N., Snowdon, L.R., Villarreal-Barajas, J.E., Sonei, R.,
IEA, 2021. Global Hydrogen Review 2021. Available online at. https://www.iea.org/re Paredes Gutiérrez, L.C., Huang, H., Stopford, A., Oldenburg, T.B.P., Zhao, J.,
ports/global-hydrogen-review-2021. Weerawardhena, P., Nightingale, M., Mayer, B., Pedersen, J.H., di Primio, R., 2019.
Ikuta, D., Ohtani, E., Sano-Furukawa, A., Shibazaki, Y., Terasaki, H., Yuan, L., Hattori, T., The dating of petroleum fluid residence time in subsurface reservoirs. Part 1: A
2019. Interstitial hydrogen atoms in face-centered cubic iron in the Earth’s core. Sci. radiolysis-based geochemical toolbox. Geochim. Cosmochim. Acta 261, 305–326.
Rep. 9, 7108. https://doi.org/10.1038/s41598-019-43601-z. Lash, G.G., Lash, E.P., 2014. Early history of the natural gas industry, Fredonia, New
Ito, T., Nagamine, K., Yamamoto, K., Adachi, M., Kawabe, I., 1999. Preseismic hydrogen York. In: Search and Discovery Article #70000. Available at: https://www.search
gas anomalies caused by stress-corrosion process preceding earthquakes. Geophys. anddiscovery.com/pdfz/documents/2014/70168lash/ndx_lash.pdf.html (accessed
Res. Lett. 26, 2009–2012. 15 December 2021).
Jeffrey, A.W.A., Kaplan, I.R., 1988. Hydrocarbons and inorganic gases in the Gravberg-I Lavrushin, V.Yu, Aydarkozhina, A.S., Pokrovsky, B.G., Prasolov, E.M., Potapov, E.G.,
well, Siljan Ring, Sweden. In: Schoell, M. (Ed.), Origins of Methane in the Earth, Ermakov, A.V., 2020. Nitrogen and carbon isotope composition of gases in high-
Chem. Geol., 71, pp. 237–255. pCO2 waters in the North Caucasus. Geochem. Int. 58, 1262–1277. https://doi.org/
Jüntgen, H., van Heek, K.H., 1979. An update of German non-isothermal coal pyrolysis 10.1134/S0016702920110087.
work. Fuel Process. Technol. 2, 261–293. Lazar, C., Cooperdock, E.H.G., Seymour, B.H.T., 2021. A continental forearc serpentinite
Kalmychkov, G.V., Hachikubo, A., Pokrovsky, B.G., Minami, H., Yamashita, S., diapir with deep origins: elemental signatures of a mantle wedge protolith and slab-
Khlystov, O.M., 2020. Methane with abnormally high δ13C and δD values from the derived fluids at New Idria, California. Lithos 398–399, 106252. https://doi.org/10
coastal hot springs in Lake Baikal. Lithol. Miner. Resour. 55, 439–444. .1016/j.lithos.2021.106252.
Kameda, J., Saruwatari, K., Tanaka, H., Tsunomori, F., 2004. Mechanisms of hydrogen Lecher, A., Chuang, P., Singleton, M., Paytan, A., 2017. Source of methane to an Arctic
generation during the mechanochemical treatment of biotite within D2O media. lake in Alaska: an isotopic investigation. J. Geophys. Res. Biogeosci. 122, 753–766.
Earth Planets Space 56, 1241–1245. https://doi.org/10.1186/BF03353346. Lefeuvre, N., Truche, L., Donzé, F.V., Ducoux, M., Barré, G., Fakoury, R.A., Calassou, S.,
Katz, B.J., 2011. Microbial processes and natural gas accumulations. Open Geol. J. 5, Gaucher, E.C., 2021. Native H2 exploration in the Western Pyrenean Foothills.
75–83. Geochem. Geophys. Geosyst. 22, 8.
Kawagucci, S., Toki, T., Ishibashi, J., Takai, K., Ito, M., Oomori, T., Gamo, T., 2010. Lehmann, L., Baerns, M., 1992. Mechanistic aspects of the oxidative coupling of methane
Isotopic variation of molecular hydrogen in 20◦ -375◦ C hydrothermal fluids as over a NaOH/CaO catalyst. Catal. Today 13, 265–272. https://doi.org/10.1016/
detected by a new analytical method. J. Geophys. Res. 115, G03021. https://doi. 0920-5861(92)80150-L.
org/10.1029/2009JG001203. Leila, M., Lévy, D., Battani, A., Piccardi, L., Šegvić, B., Badurina, L., Pasquet, G.,
Kelemen, P.B., Matter, J., 2008. In situ carbonation of peridotite for CO2 storage. Combaudon, V., Moretti, I., 2021. Origin of continuous hydrogen flux in gas
Proceed. Nat. Acad. Sci. 105, 17295–17300. manifestations at the Larderello geothermal field, Central Italy. Chem. Geol. 120564
Kelley, D.S., Früh-Green, G.L., 1999. Abiogenic methane in deep-seated mid-ocean ridge https://doi.org/10.1016/j.chemgeo.2021.120564.
environments; insights from stable isotope analyses. J. Geophys. Res. B 104, Leila, M., Loiseau, K., Moretti, I., 2022. Controls on generation and accumulation of
10,439–10,460. blended gases (CH4/H2/He) in the Neoproterozoic Amadeus Basin, Australia. Mar.
Kenney, J.F., Kutcherov, V.A., Bendeliani, N.A., Alekseev, V.A., 2002. The evolution of Pet. Geol. 140, 105643 https://doi.org/10.1016/j.marpetgeo.2022.105643.
multicomponent systems at high pressures: VI. The thermodynamic stability of the Levshounova, S.P., 1991. Hydrogen in petroleum geochemistry. Terra Nova 3, 579–585.
hydrogen-carbon system: the genesis of hydrocarbons and the origin of petroleum. Lewan, M.D., Ulmishek, G.F., Harrison, W., Schreiner, F., 1991. Gamma-Co-60-
Proc. National Acad. Sci. 99, 10976–10981. irradiation of organic matter in the phosphoria retort shale. Geochim. Cosmochim.
King, G.E., 2012. Hydraulic fracturing 101: what every representative, environmentalist, Acta 55, 1051–1063.
regulator, reporter, investor, University researcher, neighbor and engineer should Li Vigni, L., Daskalopoulou, K., Calabrese, S., Parello, F., D’Alessandro, W., 2021.
know about estimating frac risk and improving frac performance in unconventional Geochemical characterisation of the alkaline and hyperalkaline groundwater in the
gas and oil wells. In: Paper presented at the SPE Hydraulic Fracturing Technology Othrys Ophiolite Massif, central Greece. Ital. J. Geosci. 140, 42–56.
Conference, The Woodlands, Texas, USA, February 2012. https://doi.org/10.2118/ Li, L., Cartigny, P., Ader, M., 2009. Kinetic nitrogen isotope fractionation associated with
152596-MS. thermal decomposition of NH3: experimental results and potential applications to
Kingsley, R.H., Schilling, J.-G., Dixon, J.E., Swart, P., Poreda, R., Simons, K., 2002. D/H trace the origin of N2 in natural gas and hydrothermal systems. Geochim.
ratios in basalt glasses from the Salas y Gomez mantle plume interacting with the Cosmochim. Acta 73, 6282–6297. https://doi.org/10.1016/j.gca.2009.07.016.
East Pacific Rise: Water from old D-rich recycled crust or primordial water from the Li, X., Krooss, B.M., Weniger, P., Littke, R., 2015. Liberation of molecular hydrogen (H2)
lower mantle? Geochem. Geophys. Geosyst. 3, 1–26. https://doi.org/10.1029/ and methane (CH4) during non-isothermal pyrolysis of shales and coals: Systematics
2001GC000199. and quantification. Int. J. Coal Geol. 137, 152–164. https://doi.org/10.1016/j.
coal.2014.11.011.

23
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Li, X., Krooss, B.M., Weniger, P., Littke, R., 2017. Molecular hydrogen (H2) and light Meincke, W., 1967. Zur herkunft des wasserstoffs in tiefenproben. Z. Angew. Geol. 13,
hydrocarbon gases generation from marine and lacustrine source rocks during 346–348 (in German).
closed-system laboratory pyrolysis experiments. J. Analyt. Appl. Pyrol. 126, Melton, C.E., Giardini, A.A., 1974. The composition and significance of gas released from
275–287. https://doi.org/10.1016/j.jaap.2017.05.019. natural diamonds from Africa and Brazil. Am. Mineral. 59, 775–782.
Lilley, M.D., Butterfleld, D.A., Olson, E.J., Lupton, J.E., Macko, S.A., McDuff, R.E., 1993. Ménez, B., 2020. Abiotic hydrogen and methane: fuels for life. Elements 16, 39–46.
Anomalous CH4 and NH+ 4 concentrations at an unsedimented mid-ocean-ridge Meng, Q., Sun, Y., Tong, J., Fu, Q., Zhu, J., Zhu, D., Jin, Z., 2015. Distribution and
hydrothermal system. Nature 364, 45–47. geochemical characteristics of hydrogen in natural gases from the Jiyang Depression,
Lin, L.-H., Hall, J., Lippmann-Pipke, J., Ward, J.A., Sherwood Lollar, B., de Flaun, M., Eastern China. Acta Geol. Sin. 89, 1616–1624.
Rothmel, M., Moser, D., Gihring, T.M., 2005a. Radiolytic H2 in continental crust: Merlivat, L., Pineau, F., Javoy, M., 1987. Hydrothermal vent waters at 13◦ N on the East
nuclear power for deep subsurface microbial communities. Geochem. Geophys. Pacific Rise: isotopic composition and gas concentration. Earth Planet. Sci. Lett. 84,
Geosyst. 6, Q07003. https://doi.org/10.1029/2004GC000907. 100–108.
Lin, L.-H., Slater, G.F., Sherwood Lollar, B., Lacrampe-Couloume, G., Onstott, T.C., Mermelstein, L.D., Zeikus, J.G., 1998. Anaerobic nonmethanogenic extremophiles. In:
2005b. The yield and isotopic composition of radiolytic H2, a potential energy source Horikoshi, K., Grant, W.D. (Eds.), Extremophiles: Microbial Life in Extreme
for the deep subsurface biosphere. Geochim. Cosmochim. Acta 69, 893–903. https:// Environment. Wiley-Liss, New York, pp. 255–285.
doi.org/10.1016/j.gca.2004.07.032. Messer, B., Oprea, V., Beaulieu, K., Wright, A., 2008. Role of nascent hydrogen in
Link, W.K., 1952. Significance of oil and gas seeps in world oil exploration. AAPG Bull. refinery corrosion. In: Paper presented at the CORROSION 2008, New Orleans,
36, 1505–1540. Louisiana, March 2008. Paper Number NACE-08549.
Liu, C., Liu, J., Sun, P., Zhang, L., Li, H., Zheng, S., Ge, Y., 2013. Geochemical features of Milesi, V., Guyot, F., Brunet, F., Richard, L., Recham, N., Benedetti, M., Dairou, J.,
natural gas in the Qaidam Basin, NW China. J. Pet. Sci. Eng. 110, 85–93. Prinzhofer, A., 2015. Formation of CO2, H2 and condensed carbon from siderite
Liu, S., Tan, F., Huo, T., Tang, S., Zhao, W., Chao, H., 2020. Origin of the hydrate bound dissolution in the 200-300◦ C range and at 50 MPa. Geochim. Cosmochim. Acta 154,
gases in the Juhugeng Sag, Muli Basin, Tibetan Plateau. Int. J. Coal Sci. Technol. 7, 201–211. https://doi.org/10.1016/j.gca.2015.01.015.
43–57. https://doi.org/10.1007/s40789-019-00283-2. Milesi, V., Prinzhofer, A., Guyot, F., Benedetti, M., Rodrigues, R., 2016. Contribution of
Loewen, M.W., Graham, D.W., Bindeman, I.N., Lupton, J.E., Garcia, M.O., 2019. siderite–water interaction for the unconventional generation of hydrocarbon gases in
Hydrogen isotopes in high 3He/4He submarine basalts: primordial vs. recycled water the Solimões basin, north-west Brazil. Mar. Pet. Geol. 71, 168–182. https://doi.org/
and the veil of mantle enrichment. Earth Planet. Sci. Lett. 508, 62–73. https://doi. 10.1016/j.marpetgeo.2015.12.022.
org/10.1016/j.epsl.2018.12.012. Milkov, A.V., 2004. Global estimates of hydrate-bound gas in marine sediments: how
Löffler, M., Schrader, M., Lüders, K., Werban, U., Hornbruch, G., Dahmke, A., Vogt, C., much is really out there? Earth Sci. Rev. 66, 183–197. https://doi.org/10.1016/j.
Richnow, H.H., 2022. Stable hydrogen isotope fractionation of hydrogen in a field earscirev.2003.11.002.
injection experiment: simulation of a gaseous H2 leakage. ACS Earth Space Chem. 6, Milkov, A.V., 2005. Molecular and stable isotope composition of natural gas hydrates: a
631–641. https://doi.org/10.1021/acsearthspacechem.1c00254. revised global dataset and basic interpretations in the context of geological settings.
Lu, Z., Zhu, Y., Zhang, Y., Wen, H., Li, Y., Liu, C., 2011. Gas hydrate occurrences in the Org. Geochem. 36, 681–702.
Qilian Mountain permafrost, Qinghai Province, China. Cold Reg. Sci. Technol. 66, Milkov, A.V., 2010. Methanogenic biodegradation of petroleum in the West Siberian
93–104. basin (Russia): significance for formation of giant Cenomanian gas pools. AAPG Bull.
Lunsford, J.H., 1990. The catalytic conversion of methane to higher hydrocarbons. Catal. 94, 1485–1541.
Today 6, 235–259. https://doi.org/10.1016/0920-5861(90)85004-8. Milkov, A.V., 2011. Worldwide distribution and significance of secondary microbial
Luo, Y.-H., Sternberg, L., Suda, S., Kumazawa, S., Mitsui, A., 1991. Extremely low D/H methane formed during petroleum biodegradation in conventional reservoirs. Org.
ratios of photoproduced hydrogen by cyanobacteria. Plant Cell Physiol. 32, Geochem. 42, 184–207.
897–900. Milkov, A.V., 2018. Secondary microbial gas. In: Wilkes, H. (Ed.), Hydrocarbons, Oils
Luong, L.D., Obzhirov, A.I., Hoang, N., Shakirov, R.B., Anh, L.D., Syrbu, N.S., Tuan, D. and Lipids: Diversity, Origin, Chemistry and Fate. Handbook of Hydrocarbon and
M., Tao, N.V., Huong, T.T., Cuong, D.H., Kholmogorov, A.O., Binh, P.V., Lipid Microbiology. Springer, Cham, pp. 1–10. https://doi.org/10.1007/978-3-319-
Mishukov, O.V., Eskova, A.I., 2021. Distribution of gases in bottom sediments of the 54529-5_22-1.
Southwestern Sub-Basin South China Sea (Bien Dong). Russ. J. Pacif. Geol. 15, Milkov, A.V., 2021. New approaches to distinguish shale-sourced and coal-sourced gases
144–154. in petroleum systems. Org. Geochem. 158, 104271 https://doi.org/10.1016/j.
Lyon, G., Hulston, J., 1984. Carbon and hydrogen isotopic compositions of New Zealand orggeochem.2021.104271.
geothermal gases. Geochim. Cosmochim. Acta 48, 1161–1171. https://doi.org/ Milkov, A.V., Etiope, G., 2018. Revised genetic diagrams for natural gases based on a
10.1016/0016-7037(84)90052-8. global dataset of >20,000 samples. Org. Geochem. 125, 109–120. https://doi.org/
Magoon, L.B., Dow, W.G., 1994. The petroleum system. In: Magoon, L.B., Dow, W.G. 10.1016/j.orggeochem.2018.09.002.
(Eds.), The Petroleum System - From Source to Trap, AAPG Memoir, 60, pp. 3–24. Milkov, A.V., Navidi, W.C., 2020. Randomness, serendipity and luck in petroleum
Mao, H.-K., Hu, Q., Yang, L., Liu, J., Kim, D.Y., Meng, Y., Zhang, L., Prakapenka, V.B., exploration. AAPG Bull. 104, 145–176.
Yang, W., Mao, W.L., 2017. When water meets iron at Earth’s core–mantle Milkov, A.V., Claypool, G.E., Lee, Y.-J., Torres, M.E., Borowski, W.S., Tomaru, H.,
boundary. Nation. Sci. Rev. 4, 870–878. https://doi.org/10.1093/nsr/nwx109. Sassen, R., Long, P.E., ODP Leg 204 Scientific Party, 2004. Ethane enrichment and
Mariner, R.H., Venezky, D.Y., Hurwitz, S., 2006. Chemical and isotopic database of water propane depletion in subsurface gases indicate gas hydrate occurrence in marine
and gas from hydrothermal systems with an emphasis for the western United States. sediments at southern Hydrate Ridge offshore Oregon. Org. Geochem. 35,
In: United States Geological Survey, Data Series, 169. Available online at. https 1067–1080.
://pubs.usgs.gov/ds/2005/169/. Milkov, A.V., Goebel, E., Dzou, L., Fisher, D.A., Kutch, A., McCaslin, N., Bergman, D.,
Marques, J.M., Etiope, G., Neves, M.O., Carreira, P.M., Rocha, C., Vance, S.D., 2007. Compartmentalization and time-lapse geochemical reservoir surveillance of
Christensen, L., Miller, A.Z., Suzuki, S., 2018. Linking serpentinization, the Horn Mountain oil field, deep-water Gulf of Mexico. AAPG Bull. 91, 847–876.
hyperalkaline mineral waters and abiotic methane production in continental Milkov, A.V., Faiz, M., Etiope, G., 2020a. Geochemistry of shale gases from around the
peridotites: an integrated hydrogeological-bio-geochemical model from the Cabeço world: Composition, origins, isotope reversals and rollovers, and implications for the
de Vide CH4-rich aquifer (Portugal). Appl. Geochem. 96, 287–301. https://doi.org/ exploration of shale plays. Org. Geochem. 143, 103997 https://doi.org/10.1016/j.
10.1016/j.apgeochem.2018.07.011. orggeochem.2020.103997.
Marty, B., Gunnlaugsson, E., Jambon, A., Oskarsson, N., Ozima, M., Pineau, F., Milkov, A.V., Schwietzke, S., Allen, G., Sherwood, O.A., Etiope, G., 2020b. Using global
Torssander, P., 1991. Gas geochemistry of geothermal fluids, the Hengill area, isotopic data to constrain the role of shale gas production in recent increases in
southwest rift zone of Iceland. Chem. Geol. 91, 207–225. atmospheric methane. Sci. Rep. 10, 4199. https://doi.org/10.1038/s41598-020-
Mayhew, L., Ellison, E., McCollom, T., Trainor, T.P., Templeton, A.S., 2013. Hydrogen 61035-w.
generation from low-temperature water–rock reactions. Nat. Geosci. 6, 478–484. Miller, H.M., Mayhew, L.E., Ellison, E.T., Kelemen, P., Kubo, M., Templeton, A.S., 2017.
https://doi.org/10.1038/ngeo1825. Low temperature hydrogen production during experimental hydration of partially-
McCarthy Jr., J.H., Cunningham, K.I., Roberts, A.A., Dietrich, J.A., 1986. Soil gas studies serpentinized dunite. Geochim. Cosmochim. Acta 209, 161–183.
around hydrogen-rich natural gas wells in northern Kansas. In: United States Moeck, I.S., 2014. Catalog of geothermal play types based on geologic controls. Renew.
Department of the Interior Geological Survey, Report 86-461, 21 p. Sustain. Energy Rev. 37, 867–882.
McCollom, T.M., Seewald, J.S., 2006. Carbon isotope composition of organic compounds Molchanov, V., 1981. Generation of Hydrogen in Lithogenesis. Nauka, Novosibirsk (in
produced by abiotic synthesis under hydrothermal conditions. Earth Planet. Sci. Lett. Russian).
243, 74–84. Monnin, C., Quéméneur, M., Price, R., Jeanpert, J., Maurizot, P., Boulart, C., Donval, J.
McCollom, T.M., Seewald, J.S., 2013. Serpentinites, hydrogen, and life. Elements 9, P., Pelletier, B., 2021. The chemistry of hyperalkaline springs in serpentinizing
129–134. environments: 1. the composition of free gases in New Caledonia compared to other
McCollom, T.M., Klein, F., Ramba, M., 2022. Hydrogen generation from serpentinization springs worldwide. J. Geophys. Res. Biogeosci. 126 e2021JG006243.
of iron-rich olivine on Mars, icy moons, and other planetary bodies. Icarus 372, Moody, J.B., 1976. Serpentinization: a review. Lithos 9, 125–138.
114754. https://doi.org/10.1016/j.icarus.2021.114754. Morita, R.Y., 2000. Is H2 the universal energy source for long-term survival? Microb.
McDermott, J.M., Sylva, S.P., Ono, S., German, C.R., Seewald, J.S., 2018. Geochemistry Ecol. 38, 307–320.
of fluids from Earth’s deepest ridge-crest hot-springs: Piccard hydrothermal field, Mørkved, P.T., Thiessen, O., Johansen, H., Dugstad, A., 2009. H2 in well casings at the
Mid-Cayman Rise. Geochim. Cosmochim. Acta 228, 95–118. Troll A platform: is corrosion a major gas source?. In: Abstracts the 24th
McDowell, B.P., Milkov, A.V., Anderson, D.S., 2017. The helium system: a modification International Meeting on Organic Geochemistry, September 6-11, 2009, Bremen,
of the petroleum system for inert gases. In: Search and Discovery Article #42098. Germany, p. 472.
Available at. https://www.searchanddiscovery.com/pdfz/documents/2017/42 Morrill, P.L., Kuenen, J.G., Johnson, O.J., Suzuki, S., Rietze, A., Sessions, A.L., Fogel, M.
098mcdowell/ndx_mcdowell.pdf.html (accessed 15 December 2021). L., Nealson, K.H., 2013. Geochemistry and geobiology of a present-day

24
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

serpentinization site in California: The Cedars. Geochim. Cosmochim. Acta 109, Proskurowski, G., Lilley, M.D., Olson, E.J., 2008a. Stable isotopic evidence in support of
222–240. https://doi.org/10.1016/j.gca.2013.01.043. active microbial methane cycling in low-temperature diffuse flow vents at 9◦ 50′ N
Murray, J., Clement, A., Frita, B., Schmittbuhl, J., Boardmann, V., Fleury, J.M., 2020. East Pacific Rise. Geochim. Cosmochim. Acta 72, 2005–2023. https://doi.org/
Abiotic hydrogen generation from biotite-rich granite: a case study of the Soultz- 10.1016/j.gca.2008.01.025.
sous-Forets geothermal site. France. Appl. Geochem. 119, 104631 https://doi.org/ Proskurowski, G., Lilley, M.D., Seewald, J.S., Früh-Green, G.L., Olson, E.J., Lupton, J.E.,
10.1016/j.apgeochem.2020.104631. Sylva, S.P., Kelley, D.S., 2008b. Abiogenic hydrocarbon production at Lost City
Nandi, R., Sengupta, S., 1998. Microbial production of hydrogen: an overview. Crit. Rev. hydrothermal field. Science 319, 604–607.
Microbiol. 24, 61–84. Pudovskis, V., 1959. Meda No. 1 Geological Completion Report. Available online at.
Naumenko-Dézes, M., Kloppmann, W., Blessing, M., Bondu, R., Gaucher, E.C., Mayer, B., https://wapims.dmp.wa.gov.au/WAPIMS/Search/WellDetails#.
2022. Natural gas of radiolytic origin: an overlooked component of shale gas. Proc. Rezaee, R., 2021. Assessment of natural hydrogen systems in Western Australia. Int. J.
Natl. Acad. Sci. 119, e2114720119 https://doi.org/10.1073/pnas.2114720119. Hydrog. Energy 46, 33068–33077. https://doi.org/10.1016/j.
Neal, C., Stranger, G., 1983. Hydrogen generation from mantle source rocks in Oman. ijhydene.2021.07.149.
Earth Planet. Sci. Lett. 66, 315–320. https://doi.org/10.1016/0012-821X(83)90144- Ricci, A., Kleine, B.I., Fiebig, J., Gunnarsson-Robin, J., Kamunya, K.M., Mountain, B.,
9. Stefánsson, A., 2022. Equilibrium and kinetic controls on molecular hydrogen
Nisbet, E.G., Sleep, N.H., 2001. The habitat and nature of early life. Nature 409, abundance and hydrogen isotope fractionation in hydrothermal fluids. Earth Planet.
1083–1091. Sci. Lett. 579, 117338 https://doi.org/10.1016/j.epsl.2021.117338.
Nivin, V.A., 2019. Occurrence forms, composition, distribution, origin and potential Richet, P., Bottinga, Y., Javoy, M., 1977. A review of hydrogen, carbon, nitrogen,
hazard of natural hydrogen–hydrocarbon gases in ore deposits of the Khibiny and oxygen, sulphur, and chlorine stable isotope fractionation among gaseous molecules.
Lovozero Massifs: a review. Minerals 9, 535. https://doi.org/10.3390/min9090535. Annu. Rev. Earth Planet. Sci. 5, 65–110.
Nothaft, D.B., Templeton, A.S., Boyd, E.S., Matter, J.M., Stute, M., Paukert Vankeuren, A. Röling, W.F.M., Head, I.M., Larter, S.R., 2003. The microbiology of hydrocarbon
N., The Oman Drilling Project Science Team, 2021. Aqueous geochemical and degradation in subsurface petroleum reservoirs: perspectives and prospects. Res.
microbial variation across discrete depth intervals in a peridotite aquifer assessed Microbiol. 154, 321–328.
using a packer system in the Samail Ophiolite, Oman. J. Geophys. Res. Biogeosci. Rӧper, M., 1983. Fischer-Tropsch synthesis. In: Keim, W. (Ed.), Catalysis in C1 chemistry.
126 https://doi.org/10.1029/2021JG006319 e2021JG006319. D. Reidel, Dordrecht, Holland, pp. 41–88.
Nuttall, W.J., Bakenne, A.T., 2020. Fossil Fuel Hydrogen: Technical, Economic and Rusakov, O., 2020. A global inventory of concentration measurements of free and
Environmental Potential. Springer, 138 p. dissolved in underground waters molecular hydrogen in the Earth’s crust on land.
Ohara, Y., Reagan, M.K., Fujikura, K., Watanabe, H., Michibayashi, K., Ishii, T., Stern, R. Geofizich. Zhurn. 42, 59–99 (in Russian). 10.24028/gzh.0203-3100.v42i6.2020
J., Pujana, I., Martinez, F., Girard, G., Ribeiro, J., Brounce, M., Komori, N., Kino, M., .222284.
2012. A serpentinite-hosted ecosystem in the Southern Mariana Forearc. Proc. Natl. Russell, M.J., Hall, A.J., Martin, W., 2010. Serpentinization as a source of energy at the
Acad. Sci. 109, 2831–2835. origin of life. Geobiology 8, 355–437.
Ostertag-Henning, C., Scheeder, G., 2009. The role of molecular hydrogen for the process Sabatier, P., Senderens, J.B., 1902. New synthesis of methane. Compt. Rend. 134,
of thermochemical sulfate reduction. In: Abstracts the 24th International Meeting on 514–516.
Organic Geochemistry, September 6-11, 2009, Bremen, Germany, p. 473. Sader, J.A., Harrison, A.L., McClenaghan, M.B., Hamilton, S.M., Clark, I.D., Sherwood
Oze, C., Sharma, M., 2005. Have olivine, will gas: serpentinization and the abiogenic Lollar, B., Leybourne, M.I., 2021. Generation of high-pH groundwaters and H2 gas by
production of methane on Mars. Geophys. Res. Lett. 32, L10203. groundwater–kimberlite interaction, northeastern Ontario, Canada. The Canad.
Oze, C., Sharma, M., 2007. Serpentinization and the inorganic synthesis of H2 in Mineral. 59, 1261–1276. https://doi.org/10.3749/canmin.2000048.
planetary surfaces. Icarus 186, 557–561. Sano, Y., Urabe, A., Wakita, H., Chiba, H., Sakai, H., 1985. Chemical and isotopic
Parnell, J., Blamey, N., 2017a. Global hydrogen reservoirs in basement and basins. compositions of gases in geothermal fluids in Iceland. Geochem. J. 19, 135–148.
Geochem. Trans. 18, 1–8. https://doi.org/10.1186/s12932-017-0041-4. Saruwatari, K., Kamed, J., Tanaka, H., 2004. Generation of hydrogen ions and hydrogen
Parnell, J., Blamey, N., 2017b. Hydrogen from radiolysis of aqueous fluid inclusions gas in quartz-water crushing experiments: an example of chemical processes in
during diagenesis. Minerals 7, 130. https://doi.org/10.3390/min7080130. active faults. Phys. Chem. Miner. 31, 176–182.
Pastina, B., LaVerne, J., 2001. Effect of molecular hydrogen on hydrogen peroxide in Sassen, R., Sweet, S.T., DeFreitas, D.A., Milkov, A.V., 2000. Exclusion of 2-methylbutane
water radiolysis. J. Phys. Chem. A 105, 9316–9322. (isopentane) during crystallization of structure II gas hydrate in sea-floor sediment,
Pepper, A.S., Corvi, P.J., 1995. Simple kinetic models of petroleum formation. Part III: Gulf of Mexico. Org. Geochem. 31, 1257–1262.
modelling an open system. Mar. Pet. Geol. 12, 417–452. Sauvage, J.F., Flinders, A., Spivack, A.J., Pockalny, R., Dunlea, A.G., Anderson, C.H.,
Perevozchikov, G.V., 2011. Hydrogen in depths of Kyzylkums. Prosp. Protect. Miner. Smith, D.C., Murray, R.W., D’Hondt, S., 2021. The contribution of water radiolysis to
Resourc. 2, 35–39 (in Russian). marine sedimentary life. Nat. Commun. 12, 1297. https://doi.org/10.1038/s41467-
Pester, N.J., Conrad, M.E., Knauss, K.G., DePaolo, D.J., 2018. Kinetics of D/H isotope 021-21218-z.
fractionation between molecular hydrogen and water. Geochim. Cosmochim. Acta Schimmelmann, A., Sauer, P.E., 2017. Hydrogen isotopes. In: White, W.M. (Ed.),
242, 191–212. https://doi.org/10.1016/j.gca.2018.09.015. Encyclopedia of Geochemistry. Springer. https://doi.org/10.1007/978-3-319-
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide. Cambridge 39193-9_326-1.
University Press, Cambridge, U.K. Schrenk, M.O., Brazelton, W.J., Lang, S.Q., 2013. Serpentinization, carbon, and deep life.
Petersen, H.I., Hillock, P., Milner, S., Pendlebury, M., Scarlett, D., 2019. Monitoring gas Rev. Mineral. Geochem. 75, 575–606. https://doi.org/10.2138/rmg.2013.75.18.
distribution and origin in the Culzean field, UK Central North Sea, using data from a Schulz, H., 1999. Short history and present trends of Fischer–Tropsch synthesis. Appl.
continuous isotope logging tool and IsoTube and test samples. J. Pet. Geol. 42, Catal. A Gen. 186, 3–12.
435–450. Schwartz, E., Fritsch, J., Friedrich, B., 2013. H2-metabolizing prokaryotes. In:
Pieterse, G., Krol, M.C., Batenburg, A.M., Steele, L.P., Krummel, P.B., Langenfelds, R.L., Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K.H., Stackebrandt, E. (Eds.), The
Röckmann, T., 2011. Global modelling of H2 mixing ratios and isotopic compositions Prokaryotes. Springer, New York, NY, pp. 119–199. https://doi.org/10.1007/0-387-
with the TM5 model. Atmos. Chem. Phys. 11, 7001–7026. https://doi.org/10.5194/ 30742-7_17.
acp-11-7001-2011. Seewald, J.S., 2001. Aqueous geochemistry of low molecular weight hydrocarbons at
Potter, J., Konnerup-Madsen, J., 2003. A review of the occurrence and origin of elevated temperatures and pressures: constraints from mineral buffered laboratory
abiogenic hydrocarbons in igneous rocks. In: Petford, N., McCaffrey, K.J.W. (Eds.), experiments. Geochim. Cosmochim. Acta 65, 1641–1664. https://doi.org/10.1016/
Hydrocarbons in Crystalline Rocks, 214. Geological Society of London Special S0016-7037(01)00544-0.
Publication, pp. 151–173. Sherwood Lollar, B., Voglesonger, K., Lin, L.-H., Lacrampe-Couloume, G., Telling, J.,
Potter, J., Salvi, S., Longstaffe, F.J., 2013. Abiogenic hydrocarbon isotopic signatures in Abrajano, T.A., Onstott, T.C., Pratt, L.M., 2007. Hydrogeologic controls on episodic
granitic rocks: Identifying pathways of formation. Lithos 182-183, 114–124. https:// H2 release from precambrian fractured rocks - energy for deep subsurface life on
doi.org/10.1016/j.lithos.2013.10.001. Earth and Mars. Astrobiology 7, 971–986.
Power, I.M., Wilson, S.A., Dipple, G.M., 2013. Serpentinite carbonation for CO2 Sherwood Lollar, B., Onstott, T., Lacrampe-Couloume, G., Ballentine, C.J., 2014. The
sequestration. Elements 9, 115–121. https://doi.org/10.2113/gselements.9.2.115. contribution of the Precambrian continental lithosphere to global H2 production.
Pratt, W.E., 1934. Hydrogenation and the origin of oil. In: Wrather, W.E., Lahee, F.H. Nature 516, 379–382. https://doi.org/10.1038/nature14017.
(Eds.), Problems of Petroleum Geology, AAPG Special Publication, 6, pp. 235–245. Sherwood, S.C., Dixit, V., Salomez, C., 2018. The global warming potential of near-
Preiner, M., Xavier, J.C., Sousa, F.L., Zimorski, V., Neubeck, A., Lang, S.Q., Greenwell, H. surface emitted water vapour. Environ. Res. Lett. 13, 104006 https://doi.org/
C., Kleinermanns, K., Tüysüz, H., McCollom, T.M., Holm, N.G., Martin, W.F., 2018. 10.1088/1748-9326/aae018.
Serpentinization: connecting geochemistry, ancient metabolism and industrial Sigvaldson, G.E., Elisson, G., 1968. Collection and analysis of volcanic gases at Surtsey,
hydrogenation. Life 8, 41. https://doi.org/10.3390/life8040041. Iceland. Geochim. Cosmochim. Acta 32, 797–805.
Prinzhofer, A., Tahara Cissé, C.S., Diallo, A.B., 2018. Discovery of a large accumulation Silva, R.C., Snowdon, L.R., Huang, H., Nightingale, M., Becker, V., Taylor, S., Mayer, B.,
of natural hydrogen in Bourakebougou (Mali). Int. J. Hydrog. Energy 43, Pedersen, J.H., di Primio, R., Larter, S., 2019. Radiolysis as a source of 13C depleted
19315–19326. natural gases in the geosphere. Org. Geochem. 138, 103911 https://doi.org/
Price, H., Jaeglé, L., Rice, A., Quay, P., Novelli, P.C., Gammon, R., 2007. Global budget of 10.1016/j.orggeochem.2019.103911.
molecular hydrogen and its deuterium content: constraints from ground station, Simon, L., Lécuyer, C., Martineau, F., Robert, F., 2011. Experimental study of D/H
cruise, and aircraft observations. J. Geophys. Res. 112, D22108. https://doi.org/ fractionation between water and hydrogen gas during the oxidation of Fe-bearing
10.1029/2006JD008152. silicates at high temperatures (600 ◦ C-1200 ◦ C). Centr. Europ. Geol. 54, 81–93.
Prinzhofer, A., Moretti, I., Françolin, J., Pacheco, C., D’Agostino, A., Werly, J., Rupin, F., https://doi.org/10.1556/CEuGeol.54.2011.1–2.8.
2019. Natural hydrogen continuous emission from sedimentary basins: the example Sipma, J., Henstra, A.M., Parshina, S.N., Lens, P.N.L., Lettinga, G., Stams, A.J.M., 2006.
of a Brazilian H2-emitting structure. Int. J. Hydrog. Energy 44, 5676–5685. https:// Microbial CO conversions with applications in synthesis gas purification and bio-
doi.org/10.1016/j.ijhydene.2019.01.119. desulfurization. Crit. Rev. Biotechnol. 26, 41–65.

25
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Skelton, A., Whitmarsh, R., Arghe, F., Crill, P., Koyi, H., 2005. Constraining the rate and Tut Haklıdır, F.S., Şengün, R., Aydın, H., 2021. Characterization and comparison of
extent of mantle serpentinization from seismic and petrological data: implications geothermal fluids geochemistry within the Kızıldere Geothermal Field in Turkey:
for chemosynthesis and tectonic processes. Geofluids 5, 153–164. new findings with power capacity expanding studies. Geothermics 94, 102110.
Sleep, N.H., Meibom, A., Fridriksson, Th., Coleman, R.G., Bird, D.K., 2004. H2-rich fluids https://doi.org/10.1016/j.geothermics.2021.102110.
from serpentinization: geochemical and biotic implications. Proc. Natl. Acad. Sci. UK OGA, 2021. Oil and Gas Authority Open Data. Available at: http://data-ogauthority.
101, 12818–12823. https://doi.org/10.1073/pnas.0405289101. opendata.arcgis.com/datasets?keyword=Geochemistry (accessed 15 December
Smith, N.J.P., 2002. It’s time for explorationists to take hydrogen more seriously. First 2021).
Break 20 (4), 246–253. Vacquand, C., Deville, E., Beaumont, V., Guyot, F., Sissmann, O., Pillot, D., Arcilla, C.,
Smith, N.J.P., Shephard, T.J., Styles, M.T., Williams, G.M., 2005. Hydrogen exploration: Prinzhofer, A., 2018. Reduced gas seepages in ophiolitic complexes: evidences for
a review of global hydrogen accumulations and implications for prospective areas in multiple origins of the H2-CH4-N2 gas mixtures. Geochim. Cosmochim. Acta 223,
NW Europe. In: Doré, A.G., Vining, B.A. (Eds.), Petroleum Geology: North-West 437–461.
Europe and Global Perspectives. Proceedings of the 6th Petroleum Geology Vovk, I.F., 1987. Radiolytic salt enrichment and brines in the crystalline basement of the
Conference, Geological Society, London, Petroleum Geology Conference Series, 6, East European Platform. In: Fritz, P., Frape, S.K. (Eds.), Saline Water and Gases in
pp. 349–358. https://doi.org/10.1144/0060349. Crystalline Rocks, Geological Association of Canada Special Paper, 33, pp. 197–210.
Snodgrass, J.E., Milkov, A.V., 2020. Web-based machine learning tool that determines Voytov, G.I., 1992. Chemical and carbon-isotope fluctuations in free gases (gas jets) in
the origin of natural gases. Comput. Geosci. 145, 104595. the Khibiny. Geochem. Int. 29, 14–24.
Song, H., Ou, X., Han, B., Deng, H., Zhang, W., Tian, C., Cai, C., Lu, A., Lin, Z., Chai, L., Wagner, T., Watanabe, T., Shima, S., 2018. Hydrogenotrophic methanogenesis. In:
2021. An overlooked natural hydrogen evolution pathway: Ni2+ boosting H2O Stams, A., Sousa, D. (Eds.), Biogenesis of Hydrocarbons. Handbook of Hydrocarbon
reduction by Fe(OH)2 oxidation during low-temperature serpentinization. Angew. and Lipid Microbiology. Springer, Cham. https://doi.org/10.1007/978-3-319-
Chem. Int. Ed. 60, 24054. 53114-4_3-1.
Spinks, J.W.T., Woods, R.J., 1990. An Introduction to Radiation Chemistry. John Wiley, Wakita, H., Nakamura, Y., Kita, I., Fujii, N., Notsu, K., 1980. Hydrogen release: new
New York. indicator of fault activity. Science 210, 188–190.
Stevens, T.O., McKinley, J.P., 1995. Lithoautotrophic microbial ecosystems in deep Walshe, J.L., 2006. Degassing of hydrogen from the Earth’s core and related phenomena
basalt aquifers. Science 270, 450–454. of system Earth. Geochim. Cosmochim. Acta 70, A684. https://doi.org/10.1016/j.
Sturchio, N.C., Abrajano Jr., T.A., Murowchick, J.B., Muehlenbachs, K., 1989. gca.2006.06.1490.
Serpentinization of the Acoje massif, Zambales ophiolite, Philippines: hydrogen and Wang, X.F., Liu, W.H., Xu, Y.C., Zheng, J.J., Zhang, D.W., 2006. Thermal simulation
oxygen isotope geochemistry. Tectonophysics 168, 101–107. experimental study on the effect of water on the formation of gaseous hydrocarbons
Suda, K., Ueno, Y., Yoshizaki, M., Nakamura, H., Kurokawa, K., Nishiyama, E., in organic matters. Prog. Nat. Sci. 16, 1275–1281.
Yoshino, K., Hongoh, Y., Kawachi, K., Omori, S., Yamada, K., Yoshida, N., Wang, X., Zeng, Z., Chen, J., 2009. Serpentinization of peridotites from the southern
Maruyama, S., 2014. Origin of methane in serpentinite-hosted hydrothermal Mariana forearc. Prog. Nat. Sci. 19, 1287–1295. https://doi.org/10.1016/j.
systems: The CH4–H2–H2O hydrogen isotope systematics of the Hakuba Happo hot pnsc.2009.04.004.
spring. Earth Planet. Sci. Lett. 386, 112–125. https://doi.org/10.1016/j. Wang, W., Wang, S., Ma, X., Gong, J., 2011. Recent advances in catalytic hydrogenation
epsl.2013.11.001. of carbon dioxide. Chem. Soc. Rev. 40, 3703–3727.
Suda, K., Gilbert, A., Yamada, K., Yoshida, N., Ueno, Y., 2017. Compound- and position- Wang, W., Liu, C., Liu, W., Zhang, D., 2018. Factors influencing hydrogen yield in water
specific carbon isotopic signatures of abiogenic hydrocarbons from on–land radiolysis and implications for hydrocarbon generation: a review. Arab. J. Geosci.
serpentinite–hosted Hakuba Happo hot spring in Japan. Geochim. Cosmochim. Acta 11, 542. https://doi.org/10.1007/s12517-018-3903-x.
206, 201–215. Wang, W., Liu, C., Zhang, D., Liu, W., Chen, L., Liu, W., 2019. Radioactive genesis of
Sugisaki, R., Ido, M., Takeda, H., Isobe, Y., Hayashi, Y., Nakamura, N., Satake, H., hydrogen gas under geological conditions: an experimental study. Acta Geol. Sin. 93,
Mizutani, Y., 1983. Origin of hydrogen and carbon dioxide in fault gases and its 1125–1134.
relation to fault activity. The J. Geol. 91, 239–258. https://doi.org/10.1086/ Wang, H., Zhao, W., Cai, Y., Ye, Y., Wang, X., Bi, L., Zhang, S., 2020. Irradiation caused
628769. gas generation from organic matter: evidence from the neutron irradiation
Suzuki, N., Saito, H., Hoshino, T., 2017. Hydrogen gas of organic origin in shales and experiment. In: IOP Conf. Ser. Earth Environ. Sci, 569, p. 012090. https://doi.org/
metapelites. Int. J. Coal Geol. 173, 227–236. https://doi.org/10.1016/j. 10.1088/1755-1315/569/1/012090.
coal.2017.02.014. Wang, W., Liu, C., Liu, W., Wang, X., Guo, P., Wang, J., Wang, Z., Li, Z., Zhang, D., 2022.
Szatmari, P., 1989. Petroleum formation by Fisher-Tropsch synthesis in plate tectonics. Dominant products and reactions during organic matter radiolysis: Implications for
AAPG Bull. 73, 989–998. hydrocarbon generation of uranium-rich shales. Mar. Pet. Geol. 137, 105497
Szponar, N., Brazelton, W.J., Schrenk, M.O., Bower, D.M., Steele, A., Morrill, P.L., 2013. https://doi.org/10.1016/j.marpetgeo.2021.105497.
Geochemistry of a continental site of serpentinization, the Tablelands Ophiolite, Ward, L.K., 1933. Inflammable gasses occluded in the Pre-Paleozoic rocks of South
Gros Morne National Park: a Mars analogue. Icarus 224, 286–296. https://doi.org/ Australia. Trans. Proc. R. Soc. S. Aust. 57, 42–47.
10.1016/j.icarus.2012.07.004. Warwick, N.J., Bekki, S., Nisbet, E.G., Pyle, J.A., 2004. Impact of a hydrogen economy on
Takai, K., Gamo, T., Tsunogai, U., Nakayama, N., Hirayama, H., Nealson, K.H., the stratosphere and troposphere studied in a 2-D model. Geophys. Res. Lett. 31,
Horikoshi, K., 2004. Geochemical and microbiological evidence for a hydrogen- L05107. https://doi.org/10.1029/2003GL019224.
based, hyperthermophilic subsurface lithoautotrophic microbial ecosystem Warwick, N., Griffiths, P., Keeble, J., Archibald, A., Pyle, J., Shine, K., 2022. Atmospheric
(HyperSLiME) beneath an active deep-sea hydrothermal field. Extremophiles 8, Implications of Increased Hydrogen Use. Available at: www.gov.uk/gover
269–282. https://doi.org/10.1007/s00792-004-0386-3. nment/publications/atmospheric-implications-of-increased-hydrogen-use (accessed
Taylor, B.E., Eichelberger, J.C., Westrich, H.R., 1983. Hydrogen isotopic evidence of on May 1, 2022).
rhyolitic magma degassing during shallow intrusion and eruption. Nature 306, Weatherbee, G.D., Bartholomew, C.H., 1984. Hydrogenation of CO2 on group VIII metals
541–545. - IV. Specific activities and selectivities of silica-supported Co, Fe, and Ru. J. Catal.
Thayer, T.P., 1966. Serpentinization considered as a constant-volume metasomatic 87, 352–362.
process. Am. Mineral. 51, 685–710. Welhan, J.A., 1981. Carbon and Hydrogen Gases in Hydrothermal Systems: the Search
Thiel, J., Byrne, J.M., Kappler, A., Schink, B., Pester, M., 2019. Pyrite formation from FeS for a Mantle Source. PhD Dissertation. University of California, 194 p.
and H2S is mediated through microbial redox activity. Proc. Natl. Acad. Sci. 116, Wenger, L.M., Pottorf, R.J., Macleod, G., Otten, G., Dreyfus, S., Justwan, H.K., 2009. Drill
6897–6902. https://doi.org/10.1073/pnas.1814412116. bit metamorphism: recognition and impact on show evaluation. In: Paper presented
Thiessen, O., Mørkved, P.T., Johansen, H., van Grass, G.W., 2009. Uncommon gases with at the SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana,
high hydrogen concentrations found in well annuli on Troll A platform. In: Abstracts October 2009. Paper Number: SPE-125218-MS. https://doi.org/10.2118/125218-
the 24th International Meeting on Organic Geochemistry (IMOG 2009), September MS.
6-11, 2009, Bremen, Germany, p. 92. Wersin, P., Birgersson, M., Olsson, S., Karnland, O., Snellman, M., 2008. Impact of
Timmerberg, S., Kaltschmitt, M., Finkbeiner, M., 2020. Hydrogen and hydrogen-derived corrosion-derived iron on the bentonite buffer within the KBS-3H disposal concept.
fuels through methane decomposition of natural gas - GHG emissions and costs. The Olkiluoto site as case study. In: POSIVA 2007-11 and SKB R-08-34. Posiva Oy,
Energy Convers. Manag. 7, 100043 https://doi.org/10.1016/j.ecmx.2020.100043. Olkiluoto, Finland and Swedish Nuclear Fuel and Waste Management Co (SKB),
Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence. Springer-Verlag, Stockholm, Sweden.
Berlin. Wiersberg, T., Erzinger, J., 2008. On the origin and spatial distribution of gas at
Toulhoat, H., Zgonnik, V., 2022. Chemical differentiation of planets: a core issue. The seismogenic depths of the San Andreas Fault from drill mud gas analysis. Appl.
Astrophys. J. 924, 83. https://doi.org/10.3847/1538-4357/ac300b. Geochem. 23, 1675–1690. https://doi.org/10.1016/j.apgeochem.2008.01.012.
Truche, L., Bazarkina, E.F., 2019. Natural hydrogen the fuel of the 21st century. In: E3S Wood, B.L., 1972. Metamorphosed ultramafites and associated formations near Milford
Web of Conferences 98, 03006, WRI-16. Sound, New Zealand. New Zeal. J. Geol. Geophys. 15, 88–128. https://doi.org/
Truche, L., Joubert, G., Dargent, M., Martz, P., Cathelineau, M., Rigaudier, T., Quirt, D., 10.1080/00288306.1972.10423948.
2018. Clay minerals trap hydrogen in the Earth’s crust: evidence from the Cigar Lake Wordsworth, R., Pierrehumbert, R., 2013. Hydrogen-nitrogen greenhouse warming in
uranium deposit, Athabasca. Earth Planet. Sci. Lett. 493, 186–197. https://doi.org/ Earth early atmosphere. Science 339, 64–67. https://doi.org/10.1126/
10.1016/j.epsl.2018.04.038. science.1225759.
Truche, L., McCollom, T.M., Martinez, I., 2020. Hydrogen and abiotic hydrocarbons: World Energy Council, 2019. New hydrogen economy - hope or hype? World Energy
molecules that change the world. Elements 16, 13–18. https://doi.org/10.2138/ Council, 42 p. Available at: https://www.worldenergy.org/assets/downloads
gselements.16.1.13. /WEInnovation-Insights-Brief-New-Hydrogen-Economy-Hype-or-Hope.pdf (accessed
Tsunogae, T., Dubessy, J., 2009. Ethane- and hydrogen-bearing carbonic fluid inclusions 15 December 2021).
in a high-grade metamorphic rock. J. Mineral. Petrol. Sci. 104, 324–329. https://doi.
org/10.2465/jmps.090622f.

26
A.V. Milkov Earth-Science Reviews 230 (2022) 104063

Worman, S.L., Pratson, L.F., Karson, J.A., Klein, E.M., 2016. Global rate and distribution Zgonnik, V., 2020. The occurrence and geoscience of natural hydrogen: a comprehensive
of H2 gas produced by serpentinization within oceanic lithosphere. Geophys. Res. review. Earth Sci. Rev. 203, 103140 https://doi.org/10.1016/j.
Lett. 43, 6435–6443. https://doi.org/10.1002/2016GL069066. earscirev.2020.103140.
Worman, S.L., Pratson, L.F., Karson, J.A., Schlesinger, W.H., 2020. Abiotic hydrogen (H2) Zgonnik, V., Beaumont, V., Larin, N., Pillot, D., Deville, E., 2019. Diffused flow of
sources and sinks near the Mid-Ocean Ridge (MOR) with implications for the molecular hydrogen through the Western Hajar mountains, Northern Oman. Arab. J.
subseafloor biosphere. Proc. Natl. Acad. Sci. 117, 13283–13293. https://doi.org/ Geosci. 12 (71) https://doi.org/10.1007/s12517-019-4242-2.
10.1073/pnas.2002619117. Zhang, M., Tang, Q., Cao, C., Li, W., Wang, H., Li, Z., Yu, M., Feng, P., 2017. The origin of
Yamazaki, H., Shida, S., 1960. Radiolysis and bond stability of squalene. J. Chem. Phys. Permian Pobei ultramafic complex in the northeastern Tarim craton, western China:
32, 950–951. Evidences from chemical and C-He-Ne-Ar isotopic compositions of volatiles. Chem.
Yang, X., Keppler, H., Li, Y., 2016. Molecular hydrogen in mantle minerals. Geochem. Geol. 469, 85–96. https://doi.org/10.1016/j.chemgeo.2017.06.006.
Perspect. Lett. 2, 160–168. https://doi.org/10.7185/geochemlet.1616. Zhu, D., Liu, Q., Jin, Z., Meng, Q., Hu, W., 2017. Effects of deep fluids on hydrocarbon
Ye, X., Tao, M., Yu, C., Zhang, M., 2007. Helium and neon isotopic compositions in the generation and accumulation in Chinese petroliferous basins. Acta Geol. Sin. 91,
ophiolites from the Yarlung Zangbo River, southwestern China: the information from 301–319. https://doi.org/10.1111/1755-6724.13079.
deep mantle. Sci. China Ser. D 50, 801–802. Zimmer, M., 1993. Zur geochemie von gasen in formationsfluiden, bohrspülungen und
Yedinak, E.M., 2022. The curious case of geologic hydrogen: assessing its potential as a krustengesteinen – ergebnisse aus der kontinentalen tiefbohrung. PhD thesis. Justus-
near-term clean energy source. Joule 6, 1–6. https://doi.org/10.1016/j. Liebig-Universität Gießen.
joule.2022.01.005. Zimmer, M., Erzingeer, J., 1995. On the geochemistry of gases in formation and drilling
Yuce, G., Italiano, F., D’Alessandro, W., Yalcin, T.H., Yasin, D.U., Gulbay, A.H., fluids - results from KTB. Sci. Drill. 5, 101–109.
Ozyurt, N.N., Rojay, B., Karabacak, V., Bellomo, S., Brusca, L., Yang, T., Fu, C.C., Zinger, A.S., 1962. Molecular hydrogen in gas dissolved in waters of oil-gas fields, Lower
Lai, C.W., Ozacar, A., Walia, V., 2014. Origin and interactions of fluids circulating Volga region. Geochemistry 10, 1015–1023.
over the Amik Basin (Hatay, Turkey) and relationships with the hydrologic, geologic ZoBell, C.E., 1947. Microbial transformation of molecular hydrogen in marine sediments,
and tectonic settings. Chem. Geol. 388, 23–39. https://doi.org/10.1016/j. with particular reference to petroleum. AAPG Bull. 31, 1709–1751.
chemgeo.2014.09.006.

27

You might also like