Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-17T20:15:26.937Z Has data issue: false hasContentIssue false

16 - Carbon in the Deep Biosphere

Forms, Fates, and Biogeochemical Cycling

Published online by Cambridge University Press:  03 October 2019

Beth N. Orcutt
Affiliation:
Bigelow Laboratory for Ocean Sciences, Maine
Isabelle Daniel
Affiliation:
Université Claude-Bernard Lyon I
Rajdeep Dasgupta
Affiliation:
Rice University, Houston

Summary

Building on the synthesis of carbon reservoirs in Earth's subsurface, this chapter focuses on the forms, cycling, and fate of the carbon supporting microbial life in the terrestrial and marine subsurface. As the subsurface is estimated to host a vast reservoir of life on Earth, identifying the carbon compounds that life uses for energy and growth is key to understanding ecosystem functioning in the past and at present, and also for extrapolating these findings to the search for life in the universe. This chapter highlights advances in quantifying small carbon compounds, measuring rates of carbon turnover, and the fate of carbon in the deep biosphere.

Type
Chapter
Information
Deep Carbon
Past to Present
, pp. 480 - 523
Publisher: Cambridge University Press
Print publication year: 2019
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This content is Open Access and distributed under the terms of the Creative Commons Attribution licence CC-BY-NC-SA 4.0 https://creativecommons.org/cclicenses/

16.1 Introduction

The form, fate, and biogeochemical cycling of carbon in subsurface environments impacts and reflects microbial activity and has important implications for global elemental fluxes. Photosynthetically derived organic matter (OM) is transported to a depth where it can continue to fuel life far from solar inputs. Alternative energy-yielding reactions such as the oxidation of minerals and reduced gases can fuel life in the rocky subsurface of both the ocean and continents, altering the distribution and characteristics of carbon compounds. Nonbiological reactions such as the precipitation of calcium carbonate influence the availability of dissolved inorganic carbon for lithoautotrophs and, simultaneously, the carbon cycle over geologic time. The abundances, characteristics, and distributions of carbon in the subsurface can therefore provide an integrated history of biotic and abiotic processes and a template for interpreting similar patterns from other planetary bodies.

The goal of this chapter is to compile insights from disparate environments in order to build a mechanistic understanding of the controls on carbon abundance and distribution in the subsurface. The sections below summarize what is known from the oceanic and continental subsurface, realms that are often studied separately. We synthesize commonalities across these environments, highlight what remains unknown, and propose ideas for future directions.

One challenge with working across the marine–continental divide is that the terminology used to describe organic carbon varies between the two. We will use the following terms and abbreviations: particulate organic carbon (POC), dissolved organic carbon (DOC), and dissolved inorganic carbon (DIC). Another discrepancy between communities is in the use of units, with ppm or mg/L dominating the continental literature and μM or mM in the marine literature. We will use molar units throughout for comparison’s sake. Finally, while the soil community has moved away from the terms “refractory” and “recalcitrant” OM, they are still common in the marine community. Here, these terms refer to OM that has escaped remineralization due to its inherent molecular structure, physical associations with minerals, energetically unfavorable conditions, or the lack of a specific microbial community adapted to carry out the necessary degradative processes.

16.2 Oceanic Sedimentary Subsurface

Approximately 1.68 × 1014 g of organic carbon per year are buried in marine and estuarine sediments (Reference Martiny, Bohannan, Brown, Colwell, Fuhrman and Green1). Burial of organic carbon in sediments represents a transfer of reducing equivalents from Earth’s surface to the subsurface, thereby allowing persistence of oxidized compounds such as O2 at the surface (Reference Hedges2). The rate of burial of organic carbon in marine sediments therefore has an important influence on the redox state, and thus habitability, of Earth’s surface.

Broadly, marine sediments can be divided into river and estuarine delta systems, continental shelves and slopes, and abyssal plains (Figure 16.1). Sediments may be more finely divided into provinces based on microbial community composition, grain size, OM content, and benthic communities, among other variables (Reference Schrenk, Huber and Edwards3).

Figure 16.1 Deep biosphere locations on the continents and in the ocean.

The oxidation of organic carbon in sediments is carried out by a series of heterotrophic organisms. Macrofauna have their greatest influence on the surface sediments of continental shelfs, while the role of meiofauna and microorganisms increases with depth and where oxygen is limited (Reference Rex, Etter, Morris, Crouse, McClain and Johnson4). Remineralization within anoxic sediments is dominated by microorganisms and is most prevalent at temperatures below ~80°C, constituting ~75% of Earth’s total sediment volume of 3.01 × 108 km3 (Reference LaRowe, Burwicz, Arndt, Dale and Amend5). The composition, abundance, and activity of heterotrophic microorganisms in marine sediments therefore has a strong influence on the burial rate and chemical nature of organic carbon. While these reactions are catalyzed by enzymes, they are ultimately controlled by thermodynamics. This section will briefly review the chemical and biological factors that regulate organic carbon oxidation and burial rates, as well as some of the models that can be constructed to describe and predict those rates.

The burial rate of organic carbon in marine sediments is controlled by a range of biological and geological processes, including sedimentation rate, primary productivity, biological activity, sediment organic carbon content, chemical and physical form of organic molecules, and concentrations of oxidants (electron acceptors), as described below and in several reviews and syntheses (Reference Burdige6Reference Middelburg13). These factors are interrelated: rapid sedimentation rates influence the quality of OM delivered to the sediment surface, which in turns affects oxidation rates, oxygen exposure time (OET), quantity and composition of heterotrophic microbial communities, and concentrations of potential electron acceptors.

16.2.1 Chemical Composition

OM is delivered to marine sediments from marine sources such as sinking plankton and consumers and from terrigenous sources such as plant litter and soil OM. The chemical composition of fresh biomass is relatively well constrained and consists predominately of carbohydrates, proteins, and lipids. The composition of terrestrial material transferred by fluvial or aeolian processes ranges from fresh biomass to highly degraded and altered material. Lignin phenols synthesized solely by vascular plants have long been used to track terrestrial inputs into the ocean (Reference Hedges, Keil and Benner14). Ancient and recycled petrogenic carbon can also be remobilized from the weathering of sedimentary rocks (Reference Blair, Leithold, Ford, Peeler, Holmes and Perkey15). This suite of compounds is subject to biotic and abiotic alteration en route to marine deposition, which further diversifies the range of organic compounds present. Physical processes within the catchment are a major control on the composition and reactivity of OM delivered to the ocean by rivers (Reference Galy, Peucker-Ehrenbrink and Eglinton16), with larger inputs of both recently synthesized and ancient petrogenic organic carbon delivered in regions of higher erosional rates such as small mountainous streams (Reference Blair and Aller10,Reference Hilton, Galy, Hovius, Horng and Chen17Reference Bao, Lee, Huang, Feng, Dai and Kao19) and some Arctic rivers (Reference Opsahl, Benner and Amon20Reference Feng, Vonk, van Dongen, Gustafsson, Semiletov and Dudarev23).

Within the sediments, heterotrophic organisms and abiotic processes such as condensation reactions or sulfurization can alter the chemical structures of OM. In general terms, the heterotrophic remineralizaiton of larger organic molecules under anoxic conditions proceeds by the breakdown of polymers into monomers and oligomers, followed by smaller alcohols and organic acids, and finally methane and CO2 (Figure 16.2) (Reference Blair, Carter and Boehme24,Reference Schmitz, Daniel, Deppenmeier and Gottschalk25). As a result, small organic molecules such as acetate, ethane, propane, and methane build up in the porewaters of anaerobic sediments, with additional contributions from acetogenesis and hydrogenotrophic methanogenesis (Reference Blair, Carter and Boehme24,Reference Blair, Martens and Desmarais26Reference Heuer, Pohlman, Torres, Elvert and Hinrichs29).

Figure 16.2 Anaerobic breakdown of OM by microorganisms via (a) methanogenesis and (b) sulfidogenesis.

Ultimately, the vast majority of OM produced in the upper water column is respired, with only 1% of gross primary production escaping remineralization to be buried in the deep sediments (Reference Burdige6Reference Hedges and Keil8,Reference Suess30). Some molecules survive due to chemical structures that are inherently recalcitrant, a process called “selective preservation.” The role of this pathway is disputed. While some compounds such as highly cross-linked macromolecules are inherently less bioavailable than others, microorganisms are capable of metabolizing even ancient and highly altered OM relatively quickly under favorable conditions. Molecules may become less bioavailable due to nonbiological alteration. Abiotic sulfurization of organic molecules, deamination of peptides, and condensation of nitrogen-containing heterocyclic molecules all appear to promote the preservation of organics in sediments (Reference Burdige7,Reference Abdulla, Burdige and Komada31). Random recombination of molecules or the production of altered metabolites by heterotrophic microorganisms can also rapidly convert labile organic carbon into far less reactive material (Reference Jiao, Herndl, Hansell, Benner, Kattner and Wilhelm32,Reference Lechtenfeld, Kattner, Flerus, McCallister, Schmitt-Kopplin and Koch33). Temperature can promote some of these transformations, as discussed in detail in Section 16.4.

16.2.2 Bulk Controls on OM Preservation

In locations with relatively rapid sediment accumulation rates such as river deltas and continental shelf sediments, greater OM preservation is most closely associated with higher mineral surface areas and shorter OETs (Reference Burdige7,Reference Hartnett, Keil, Hedges and Devol34). These factors have proven broadly predictive of organic carbon distributions, although they do little to reveal the underlying mechanisms of preservation, nor do they allow for predictions of future responses to changing environmental conditions.

A prevailing paradigm is that microorganisms access POC only after it has been solubilized into DOC (Reference Hee, Pease, Alperin and Martens35). Organic molecules enter cells via general uptake porins, which can only accommodate molecules in the size range of 600–1000 Da (Reference Benz and Bauer36). Organic molecules in seawater, sediment porewater, and soils that are larger than 1000 Da are, however, more bioavailable than small molecules on average (Reference Burdige and Gardner37Reference Walker, Beaupre, Guilderson, McCarthy and Druffel40), apparently because smaller molecules tend to be more extensively modified than larger molecules (Reference Kelleher and Simpson41). Therefore, microbial extracellular enzymes appear necessary for the uptake and utilization of the most bioavailable organic carbon in sediments. Consistent with this paradigm, extracellular enzyme activity has been observed in deep, old sediments, including 217,000-year-old Mediterranean sapropels (Reference Coolen and Overmann42,Reference Coolen, Cypionka, Sass, Sass and Overmann43) and Baltic Sea sediments that are up to 10,000 years old (Reference Schmidt44).

Several findings complicate the view that extracellular enzymes catalyze the rate-limiting step in biological organic carbon oxidation. Extracellular enzyme activities can outstrip the ability of sediment microbes to take up hydrolysate on timescales of days to years, leading to accumulations of apparently bioavailable low-molecular-weight DOM (Reference Robador, Bruchert, Steen and Arnosti45). Further, cells do not exclusively take up organic compounds via general uptake porins. Active transporters, for instance, use energy gradients to pass specific molecules through the cell membrane. These can be extremely large: for instance, certain TonB-dependent transporters can import intact proteins up to 69 kDa (Reference Noinaj, Guillier, Barnard and Buchanan46). Additionally, in seawater and in cow rumen, some cells are able to take up larger oligosaccharides into their periplasm, store them over extended periods, and then metabolize them when conditions are right (Reference Arnosti and Repeta47,Reference Cuskin, Lowe, Temple, Zhu, Cameron and Pudlo48). The extent to which these mechanisms are important in sediments is not known, but temporal decoupling between macromolecule hydrolysis and metabolism could have implications for the dynamics of sediment OM oxidation.

16.2.3 Sorption

It has been observed for nearly 40 years that the volume-specific quantity of mineral surface area in sediments is correlated with organic carbon content (Reference Tanoue and Handa49). The mechanism underlying this relationship is not precisely understood. Sediments tend to accumulate quantities of organic carbon that are roughly equivalent to the amount that would be required to cover minerals in an organic monolayer (Reference Mayer50,Reference Mayer51). Sedimentary organic carbon, however, exists in discrete “blebs” (Figure 16.3), so the fact that the average quantity of OM per unit of mineral surface area is roughly monolayer equivalent appears to be essentially coincidental (Reference Mayer52).

Figure 16.3 (a) Scanning transmission X-ray microscope image and (b) optical density map of the organic carbon distribution of sediments from 1.75 m below seafloor at Integrated Ocean Drilling Program Site 1231 Hole B, Peru Basin. The optical density map was generated by subtracting a pre-edge X-ray image from a post-edge X-ray image; brighter pixels correspond to higher concentrations of organic carbon. OM associated with particles is not distributed evenly over the surface.

Image courtesy of Dr. E. Estes, University of Delaware.

Several mechanisms appear responsible for the protection of OM by mineral surfaces. First, OM may be occluded between mineral grains, within minerals themselves, or even within a matrix of more recalcitrant sorbed organic compounds (Reference Knicker and Hatcher53,Reference Moore, Nunn, Goodlett and Harvey54). Encased OM represents a sterile microenvironment in which biological oxidation is impossible. Second, even when sorbed OM is physically accessible to microorganisms, sorption slows or halts the diffusion of organic compounds to cell membranes (Reference Wu and Gschwend55). Finally, sorption distorts the physical structure of extracellular enzymes, preventing them from functioning normally, while simultaneously protecting enzymes from degradation and thereby substantially extending their active lifetimes (Reference Espeland and Wetzel56,Reference Tietjen and Wetzel57). Associations with iron oxides, which include chelation, coprecipitation, and noncovalent bonding to oxide surfaces, accounts for an average of 20% of organic carbon in sediments (Reference Lalonde, Mucci, Ouellet and Gelinas58,Reference Barber, Brandes, Leri, Lalonde, Balind and Wirick59). A full understanding of the mechanisms of microbial OM oxidation in sediments requires consideration of both the interactions between organic carbon and sediment minerals and the effects of mineral surfaces on the metabolisms of microorganisms.

16.2.4 Oxygen Exposure Time

Typical marine sediments underlying oxygenated seawater contain oxic porewater near the sediment–water interface, which becomes anoxic with increasing depth due to heterotrophic organic carbon oxidation. The depth of the oxic layer can vary dramatically, from millimeters or less in rapidly accumulating, organic-rich sediments to meters in ocean gyres. The presence of oxygen enhances the remineralization of organic molecules (Reference Cowie and Hedges60Reference Keil, Hu, Tsamakis and Hedges62), and the term “oxygen exposure time” was coined to quantify the average time that sedimentary OM is exposed to “oxic” conditions, which can range from days to thousands of years (Reference Hedges and Keil8,Reference Hartnett, Keil, Hedges and Devol34,Reference Keil, Hu, Tsamakis and Hedges62,Reference Hedges, Hu, Devol, Hartnett, Tsamakis and Keil63). Organic carbon oxidation is substantially faster in oxic sediments than anoxic sediments because the greater free energy of reaction of organic carbon with oxygen allows for a denser microbial community capable of catalyzing faster oxidation and because specific reactions (e.g. the oxidation of lignin via an oxygen radical intermediate) are not possible, or are vastly slower, in the absence of molecular oxygen (Reference Kirk and Farrell64). Thus, shorter exposure times are associated with higher organic carbon burial efficiencies and the preservation of less degraded materials (Reference Burdige7,Reference Hedges and Keil8,Reference Middelburg13,Reference Hartnett, Keil, Hedges and Devol34,Reference Hedges, Hu, Devol, Hartnett, Tsamakis and Keil63,Reference Cowie, Hedges, Prahl and Delange65,Reference Cowie, Calvert, Pedersen, Schulz and von Rad66). This correlation is not absolute, however. Large provinces of ocean sediment underlying gyres are oxic to the basement, representing as much as 86 million years of OET (Reference Røy67). In such sediments, sedimentation rates are exceedingly slow and the sediments are very organic poor, and most oxidation apparently occurs directly at the sediment–water interface. Organic carbon oxidation in “rich” anoxic systems such as rapidly accumulating estuarine sediments can exceed 100 µM C day–1, primarily via sulfate reduction, compared with ~3 × 10–6 µM C day–1 in oxic gyre sediments.

16.2.5 Models of Organic Carbon Diagenesis

Due to the chemical complexity of sedimentary OM, sedimentary diagenetic models have focused on the transformation of bulk organic carbon to CO2. One common class of models assumes the following form:

r=dGdt=i=1nkiGi,(16.1)

where r is the bulk rate of CO2 production, equivalent to the rate disappearance of bulk organic carbon (G), which in turn is the sum of the oxidation rates of different carbon pools (Gi), each of which is oxidized according to a different, characteristic rate constant (ki) (Reference Westrich and Berner68). Frequently, these “multi-G” models only include two or three reactivity classes of OM: usually a fast-reacting “labile” pool, an unreactive “recalcitrant” pool, and sometimes an intermediate “semi-labile” pool. Related models include the reactivity-continuum model, which assumes an infinite number of reactivity pools (Reference Boudreau and Ruddick69,Reference Tarutis70), and that of Middelburg (Reference Middelburg71), in which a single time-dependent reactivity rate constant is assumed. These models are mathematically straightforward, but they are somewhat mechanistically disconnected from the reality of sediment OM, which is tremendously chemically complex (Reference Burdige7,Reference Hedges, Eglinton, Hatcher, Kirchman, Arnosti and Derenne72).

Recently, models that include a broader set of parameters, such as microbial biomass, enzyme substrate specificity, and temperature–rate relationships, have been successfully employed in soils and sediments (Reference Wieder, Bonan and Allison73,Reference Bradley, Amend and LaRowe74). By including a wider range of processes, these models have the capacity to both quantitatively fit bulk organic carbon concentration data and make reasonable predictions about systems’ likely responses to changing environments.

16.3 Oceanic Rocky Subsurface

Below ocean sediments, the igneous ocean crust hosts ~2% of the total volume of the ocean, making it the largest aquifer system on Earth (Reference Johnson and Pruis75). Seawater actively circulates through this aquifer and drives the transfer of heat and elements between fluids and rocks with ramifications for ocean chemistry (Reference Edmond, Measures, McDuff, Chan, Collier and Grant76Reference German, Casciotti, Dutay, Heimburger, Jenkins and Measures79) and for the thermal, physical, and geochemical structure of the crust and mantle (Reference Stein and Stein80,Reference Alt81). Microbial life is widespread in the rocky oceanic subsurface and both exploits and influences these exchanges (Reference Summit and Baross82Reference Orcutt, Sylvan, Knab and Edwards84), altering the abundance and form of carbon. Fluid flow through the rocky subsurface is ultimately driven by a source of heat such as cooling magma or hot rocks (Reference Stein and Stein80,Reference Sclater, Jaupart and Galson85). Heated fluids rise buoyantly and ultimately exit the sub-seafloor, drawing cool seawater into the crust to replace it.

The carbon characteristics of the fluids and rocks in hydrothermal systems and the igneous basement differ greatly depending on the type of host rock, the temperature of the system, and the presence or absence of sediments. Some systems are further influenced by factors such as phase separation, magma injections, seismic activity, extent of subduction, and even tides (Reference Butterfield, McDuff, Mottl, Lilley, Lupton and Massoth86Reference Seewald, Cruse and Saccocia89). Below, carbon transformations are described in some of the primary types of hydrothermal circulation systems (Figure 16.4).

Figure 16.4 The abundance and composition of organic molecules in hydrothermal fluids will reflect a complex reaction history. While chemoautotrophy and abiotic synthesis involve the reduction of inorganic carbon into organic molecules, remineralization will do the reverse. Oxidation and dehydration reactions produce smaller, more polar compounds that are generally more labile and more easily consumed by heterotrophic microorganisms. Reduction and dehydration reactions may produce larger and more apolar material that is more resistant to microbial degradation and may be sequestered in the subsurface or persist for long periods of time in the deep ocean.

16.3.1 Characteristics of Recharge Water

The chemical composition of the seawater that enters into the rocky subsurface has a strong influence on subsequent water–rock and microbial reactions. Deep seawater carries DIC in concentrations of 2.1–2.3 mM (Reference Sarmiento and Gruber90) and DOC in concentrations of ~34–48 µM (Reference Hedges2,Reference Hansell, Carlson, Repeta and Schlitzer91). DOC is composed of a complex set of molecules, some of which turnover rapidly on timescales of hours to years. The majority of DOC, however, is slow to remineralize and has the potential to be stored for millennia in the ocean’s interior (see (Reference Hansell92) for review). Refractory DOC is highly degraded, has few recognizable biomarkers, and has a 14C age of 4000–6000 years, substantially longer than the mixing time of the ocean (Reference Druffel, Williams, Bauer and Ertel93). DOC isolated from seawater and subjected to nuclear magnetic resonance and Fourier-transform ion cyclotron resonance mass spectrometry is composed primarily of carboxyl-rich aliphatic matter (Reference Hertkorn, Benner, Frommberger, Schmitt-Kopplin, Witt and Kaiser94), acylated polysaccharides (Reference Aluwihare, Repeta and Chen95), and carotenoid degradation products (Reference Arakawa, Aluwihare, Simpson, Soong, Stephens and Lane-Coplen96) (see (Reference Repeta and Hansell97) for review).

16.3.2 Axial High Temperature, Basalt Hosted

The most widely recognized hydrothermal systems are close to axial spreading centers, where new injections of magma maintain high temperatures (Figure 16.1). The host rock is mafic and fluids exit through chimney structures at temperatures that can reach >400°C (Reference Campbell, Palmer, Klinkhammer, Bowers, Edmond and Lawrence98). Exiting fluids are rich in dissolved metals that, upon mixing with cold seawater, precipitate the sulfide minerals that give them the name “black smokers.” In the water column, the hot fluids mix further with seawater, cool, reach neutral buoyancy, and spread away from the vent field. The chemical signatures from these plumes of water can be detected thousands of kilometers away from the field (Reference Conway and John99,Reference Resing, Sedwick, German, Jenkins, Moffett and Sohst100).

The majority of high-temperature vent fluids have DIC concentrations equal to or greater than deep seawater due to inputs of magmatic CO2 (Figure 16.5 and Table 16.1) (Reference McCollom, Lowell, Seewald, Metaxas and Perfit101). DIC concentrations are generally 3–30 mM, or they can be higher when fluids are impacted by phase separation, fresh inputs of magma, or sedimentary degradation (see (Reference McCollom, Lowell, Seewald, Metaxas and Perfit101) for review). Additions of magmatic CO2 are identified by δ13C isotopic signatures (–9‰ to –4‰) (Reference Pineau and Javoy102,Reference DesMarais and Moore103) that are markedly different from deep seawater DIC (–0.5‰ to 1.0‰) (Reference Kroopnick104). The lack of 14C in CO2 from some hydrothermal fluids demonstrates that the DIC carried with recharge water can be fully removed during sub-seafloor circulation in some cases (Reference Proskurowski, Lilley and Brown105). Calcium carbonate veins in basalts and gabbros have isotope values consistent with precipitation of marine DIC at the relatively low temperatures of seawater recharge (Reference Alt and Teagle106,Reference Eickmann, Bach, Rosner and Peckmann107).

Figure 16.5 Range of methane and CO2 concentrations in basalt-hosted high-temperature (black outline; Axial Volcano, Trans-Atlantic Geotraverse (TAG), 9°N East Pacific Rise, Lucky Strike), ultramafic-hosted (green diamonds; Lost City, East Summit of Von Damm, Rainbow), ridge flank (blue checkers; Juan de Fuca ridge flank, North Pond), back-arc basins (orange diagonal; PACMANUS, Mariana Arc, Okinawa Trough), and sedimented systems (gray boxes; Guaymas, Middle Valley, Okinawa Trough). Seawater composition is included for comparison. Methane concentrations at North Pond are plotted at the reported detection limit of the analysis (0.5 µM).

References are given in Table 16.1.

Table 16.1 Characteristics and carbon contents of representative oceanic sub-seafloor fluids.

System typeSeawaterSedimentary porewaterBasalt hosted, unsedimented, high temperatureBasalt hosted, unsedimented, diffuse/mixed fluidsBasalt hosted, sedimented, high temperatureRidge Flank (“warm”)Ridge Flank (“cool”)Ultramafic influenced, high temperatureUltramafic dominatedSilicic back-arc
Example systemBelow 1500 maCascadia Marginb
0–65 mbsf
Cascadia Marginb 65–189 mbsfEast Pacific Rise (9°50'N)cAxial VolcanodTAGeEast Pacific Rise diffuse (9°50'N)fAxial VolcanogGuaymas BasinhMiddle ValleyiJuan de Fuca Ridge FlankjNorth Pond BasementkVon Damm (East Summit)lRainbowmLost CitynPACMANUSoOkinawa Trough (sedimented)p
Temperatures (°C)2–5≤9~8–12275–371217–328290–32123–553–78100–31540–281643.1–3.8226350–36730–91152–358>220–320
pH (at 23°C)7.87.7–8.07.7–8.43.5–4.23.5–4.43.1–3.55.8–6.44.6–5.84.5–6.17.57.4–7.65.62.8–3.49–112.3–4.74.7–5.4
H2 (mM)0.00030.27–8.4<0.1–0.80.10–0.37BDL–0.0058<0.50.52–3.301.9–8.20.3–0.718.2–19.212.3–16.51–140.0084–0.3060.05
Dissolved iron (µM)<0.0018–515012–10651640–5170<2–27735–40017–1800.6–1.123,700–24,050<3.5 µM76–14,600
ΣCO2 (mM)2.1–2.37–2418–299.4–21950–2852.9–5.03.0–11.835–548.2–130.2–0.62.0–2.42.8016.0–24.60.0001–0.0264.4–274198–200
δ13CCO2 (‰)–0.6 to 0.4–24.4 to 25.627.9–33.6–4.2 to –3.7–13.0 to –6.9–4.2 to –2.2–9.4–34.6 to –20.7–9.7 to –1.3–0.16 to 0.670.8–0.9–3.15 to –2.5~–9–5.7 to –2.3–5.0 to –4.7
F14CCO20.7511–0.96770.0560.083–0.2330.595–0.8650.0251–0.0373
CH4 (mM)0.0003BDL–6971–2360.05–0.160.0250.12–0.160.003–0.500<0.644.2–58.83.0–22.60.001–0.030<0.00052.811.6–2.50.9–2.00.014–0.0852.4–7.1
δ13CCH4 (‰)–59.7 ± 7.1–46.7 to –41.7–34.6 to –16.8–9.5 to –8.0–43.8–55.5 to –50.8–58 to –23–15.6 to –15.3–17.7 to –15.8–13.6 to –9.3–20.8 to –7.4–41.2 to –36.1
F14CCH40.0770.0056–0.00640.0017–0.0062
Σ(C2H6–C4H8) (µM)BDL0–35 ppmvq0.5–7318 ppmvq14–3106950.841.0–2.0
δ13CC2H6,C4H8 (‰)–25.3 to –18.7–12.9 to –9.8–16.0 to –13.0
CO (µM)BDLBDL–2.0BDL5.7–17.827–92.4n.d.5.0–7.4BDL0.006–0.17
CH3SH (nM)BDL2.4–4.91211–10,000227.4–10.31.4–1.9
DOC (µM)35–45400–32001700–51008–2434–71111–211211–1818–3368–106
δ13CDOC (‰)–20 to –22–23.6 to –22.1–20.2 ± 0.4–18.6–34.5 to –24.8–26.6 to –23.9–21.0 to –10.5
F14CDOC0.444–0.7670.481r0.166–0.230q
0.186–0.204
0.352–0.472
Formate (µM)BDL<4088.236–158
Acetate (µM)BDL5–5714–89BDL–2951–35
Hydrolizable amino acids (µM)80–1605.20.043–0.0890.7–2.3

“–” is used where no reports available in the literature.

BDL = below detection limit; mbsf = neters below seafloor; TAG = Trans-Atlantic Geotraverse.

q Headspace gas concentrations in equilibrium with sediments.

r F14C of ultrafiltrated DOC (>1000 Da).

Methane concentrations in sediment-free, high-temperature axial fluids (~7–200 µM) are higher than those of seawater (0.0003 µM), but generally low when compared to sedimented or ultramafic-influenced systems (Figure 16.5 and Table 16.1). For example, vent fluid CH4 concentrations range from 7 to 213 µM from high-temperature vents from along the East Pacific Rise (Reference Von Damm and Lilley111,Reference Von Damm123,Reference Welhan, Craig and Rona147Reference Merlivat, Pineau and Javoy150), while those from along the Mid-Atlantic Ridge (MAR) range from 8 to 147 µM (Reference Campbell, Palmer, Klinkhammer, Bowers, Edmond and Lawrence98,Reference Charlou and Donval117Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119,Reference Jeanbaptiste, Charlou, Stievenard, Donval, Bougault and Mevel151Reference Lein, Grichuk, Gurvich and Bogdanov153). Concentrations can spike as a result of volcanic eruptions and due to outgassing after a dike injection (Reference Lilley, Butterfield, Lupton and Olson88,Reference Seewald, Cruse and Saccocia89,Reference Yucel and Luther113).

The majority of the DOC carried with deep seawater is destroyed during circulation through mafic hydrothermal systems. The first evidence for this removal came from a study of amino acids in the sediment-covered Guaymas Basin (Reference Blair, Martens and Desmarais26,Reference Haberstroh and Karl122). Concentrations of dissolved free amino acids in high-temperature fluids (>150°C) were below detection limits and below deep ocean concentrations, with the losses attributed to the instability of organic compounds at high temperatures (Reference Haberstroh and Karl122,Reference Martens154). The DOC content of black smoker vents on the unsedimented portions of the Juan de Fuca spreading center is less than half that of deep seawater (<17 versus 36 µM) (Reference Lang, Butterfield, Lilley, Johnson and Hedges116). Concentrations of DOC that can be isolated onto solid-phase extraction (SPE-DOC) phases are ~92% lower in unsedimented black smokers from Juan de Fuca and the MAR than in deep seawater (Reference Hawkes, Rossel, Stubbins, Butterfield, Connelly and Achterberg155). It is possible to experimentally reproduce losses of OM by heating (Reference Lin, Koch, Feseker, Ziervogel, Goldhammer and Schmidt125,Reference Hawkes, Rossel, Stubbins, Butterfield, Connelly and Achterberg155Reference Seewald, Seyfried and Thornton157), though this does not conclusively rule out alternative removal mechanisms such as sorption onto mineral surfaces or heterotrophy.

16.3.3 Axial Diffuse Vents, Basalt Hosted

Adjacent to axial, high-temperature systems, local seawater enters the crust, creating “diffuse vents.” The mixing of oxygenated seawater and reduced hydrothermal fluids results in chemical disequilibria that microorganisms can exploit for metabolic energy (Reference McCollom and Shock158). Due to mixing and conductive cooling of fluids, temperatures are often well below the upper temperature limits of life (122°C) (Reference Takai, Nakamura, Toki, Tsunogai, Miyazaki and Miyazaki159). As a result, these zones are thriving sub-seafloor microbial habitats (Reference Schrenk, Huber and Edwards3,Reference Summit and Baross82,Reference Orcutt, Sylvan, Knab and Edwards84,Reference Karl, Wirsen and Jannasch160,Reference Edwards, Becker and Colwell161). Microbial activity can alter fluid chemistry, resulting in losses of H2S and H2 and gains of CH4 relative to high-temperature fluids (Reference Von Damm and Lilley111,Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112,Reference Butterfield, Roe, Lilley, Huber, Baross and Embley115,Reference Wankel, Germanovich, Lilley, Genc, DiPerna and Bradley162).

In diffuse vents on the Juan de Fuca Ridge, DOC is elevated over local deep seawater (~47 versus 36 µM), attributed in part to sub-seafloor autotrophic production (Reference Lang, Butterfield, Lilley, Johnson and Hedges116). This DOC has a lower 14C content and a more positive δ13C value than local seawater, consistent with a contribution of chemolithoautotrophs incorporating a pre-aged carbon source such as mantle CO2 (Reference McCarthy, Beaupre, Walker, Voparil, Guilderson and Druffel121).

16.3.4 Ridge Flanks

Fluid continues to flow through the rocky subsurface far from the ridge axis, as rocks cool in the absence of new magma injections (Figure 16.1). The extent of advective flow through these “ridge flank” systems can be determined from discrepancies between modeled conductive heat loss and heat flow measurements that indicate the convective flow of water in crust that is 0–65 Ma (Reference Johnson and Pruis75,Reference Stein and Stein80). Sediment cover precludes fluid transport into and out of the crust; bare-rock seamounts are therefore the primary locations of advective transport (Reference Wheat, Mottl, Fisher, Kadko, Davis and Baker128). Even in regions with thick sedimentary layers, however, exchange of water, carbon, elements, and nutrients continues between deep sedimentary porewater and basement fluids.

Based on magnesium budgets, fluid fluxes through “cool” ridge flank systems (<45°C) are substantially larger than those through warmer systems (Reference Mottl and Wheat77,Reference Mottl, Halbach, Tunnicliffe and Hein163). Cool basement fluids (<20°C) have been accessed by Integrated Ocean Drilling Program drilling in the North Pond sedimented basin on the MAR (Reference Edwards, Becker and Colwell161). Dorado Outcrop on the Cocos Plate has also been confirmed to vigorously vent large quantities of water at temperatures of 10–20°C (Reference Wheat, Fisher, McManus, Hulme and Orcutt164). The “warm” ridge flank system on the Juan de Fuca ridge has been intensely studied for decades, including via series of Ocean Drilling Program boreholes that have been drilled perpendicular to the ridge to allow direct access to the basement (Reference Elderfield, Wheat, Mottl, Monnin and Spiro165).

DIC is substantially lower in Juan de Fuca ridge flank fluids than in seawater (0.1–0.9 versus 2.6 mmol/kg; Table 16.1), likely due to precipitation of calcium carbonate in the subsurface (Reference Walker, McCarthy, Fisher and Guilderson130,Reference Mottl, Wheat, Baker, Becker, Davis and Feely166,Reference Sansone, Mottl, Olson, Wheat and Lilley167). In contrast, fluids from the lower-temperature Dorado and North Pond systems have DIC concentrations that are similar to seawater (Reference Meyer, Jaekel, Tully, Glazer, Wheat and Lin134,Reference Wheat, Fisher, McManus, Hulme and Orcutt164). In many cases, the δ13C values of DIC are lower than that of seawater, suggesting an input from remineralization of organic carbon or CO2 trapped in basaltic vesicles (Reference Walker, McCarthy, Fisher and Guilderson130,Reference Lin, Repeta, Xu and Rappe133,Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135). The apparent 14C age of DIC is often used as a measure of fluid residence time, although this must be treated with caution, as mixing with older water masses, remobilization of calcium carbonate, input of basalt vesicle CO2, and remineralization of 14C-depleted OM can influence these signatures (Reference Lin, Repeta, Xu and Rappe133,Reference Maloszewski and Zuber168Reference Bethke and Johnson170).

Methane concentrations are low but detectable in Juan de Fuca ridge flank fluids (1–32 µmol/kg) (Reference Lin, Cowen, Olson, Amend and Lilley131,Reference Lin, Cowen, Olson, Lilley, Jungbluth and Wilson132). The isotopic signatures of methane (–58.0‰ to –22.5‰) indicate a mixture of processes, including biogenic production and oxidation (Reference Lin, Cowen, Olson, Lilley, Jungbluth and Wilson132). Methane concentrations at North Pond were below detection (Reference Meyer, Jaekel, Tully, Glazer, Wheat and Lin134).

DOC concentrations are lower than seawater in ridge flank fluids on the Juan de Fuca ridge and at North Pond (Reference Lang, Butterfield, Lilley, Johnson and Hedges116,Reference Lin, Cowen, Olson, Amend and Lilley131,Reference Lin, Repeta, Xu and Rappe133Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135). In both cases, this DOC has a lower 14C content and δ13C signatures that are more negative than those of starting seawater (Reference McCarthy, Beaupre, Walker, Voparil, Guilderson and Druffel121,Reference Lin, Repeta, Xu and Rappe133,Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135). This pattern was initially attributed to a complete removal of seawater DOC, followed by an input of chemosynthetically derived organic material (Reference McCarthy, Beaupre, Walker, Voparil, Guilderson and Druffel121). New data suggest that the isotopic signatures could instead be attributed to the selective oxidation and removal of portions of the seawater DOC pool (Reference Lin, Repeta, Xu and Rappe133,Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135). Diffusion of porewater from the sediments covering the ridge flank may also contribute some organic compounds to the fluids (Reference Lin, Amend, LaRowe, Bingham and Cowen124), as this exchange impacts the inorganic chemistry (Reference Wheat, Hulme, Fisher, Orcutt and Becker129,Reference Lin, Cowen, Olson, Amend and Lilley131,Reference Elderfield, Wheat, Mottl, Monnin and Spiro165).

16.3.5 Ultramafic Influenced

Systems hosted on ultramafic rocks undergo water–rock reactions that are distinct from those of mafic environments. Ultramafic systems can be located on spreading centers and influenced by magmatic injections, but they can also be far from the spreading center or along ultra-slow-spreading centers with little to no magmatic influence. The compositional differences between ultramafic rocks derived predominantly from Earth’s mantle and mafic rocks such as basalt and gabbro give rise to fluids with distinct chemical signatures. Fluids that have reacted peridotites are strongly enriched in H2 and CH4 and, in some cases, have drastically lower metal contents (Figure 16.5 and Table 16.1).

The earliest recognitions of an ultramafic hydrothermal signature in the ocean came from high ratios of CH4 to Mn and suspended particulate matter in the water column on the MAR (Reference Rona, Widenfalk and Bostrom171Reference Rona, Bougault, Charlou, Appriou, Nelsen and Trefry173). Subsequently, the Logatchev, Rainbow, Menez Gwen, Ashadze, and Nibelungen hydrothermal fields were identified along the MAR, with fluid chemistries that exhibit a mixture of magmatic influences such as high temperatures (200–372°C), acidic pHs (2–4 at 25°C), and high metal contents (e.g. millimolar concentrations of Fe and hundreds of micromolar concentrations of Mn), but also ultramafic influences such as millimolar concentrations of CH4 and H2 (Table 16.1; for reviews, see Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119,Reference Schrenk, Brazelton, Lang, Hazen, Jones and Baross174). Peridotite-influenced systems have since been identified on the Mid-Cayman Rise (Reference McDermott, Seewald, German and Sylva136,Reference German, Bowen, Coleman, Honig, Huber and Jakuba175) and Marianas Forearc (Reference Mottl, Halbach, Tunnicliffe and Hein163,Reference Haggerty176). The ultramafic-dominated system in the Lost City Hydrothermal field has minimal interaction with magmatic processes, resulting in lower fluid temperatures (40–91°C), alkaline pHs (9–11 at 23°C), and low metal contents (<100 nM of Fe and <50 nM of Mn) (Reference Kelley, Karson, Früh-Green, Yoerger, Shank and Butterfield140,Reference Seyfried, Pester, Tutolo and Ding143,Reference Kelley, Karson, Blackman, Früh-Green, Butterfield and Lilley177). A magmatic influence is still evident, however, in elevated the 3He content of fluids (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112). Ultramafic-dominated, low-temperature, alkaline systems are also present in the shallow waters of Prony Bay in New Caledonia, fed by meteoric water (Reference Monnin, Chavagnac, Boulart, Menez, Gerard and Gerard178), and on the Southern Mariana Forearc at the Shinkai Seep Field (Reference Ohara, Reagan, Fujikura, Watanabe, Michibayashi and Ishii179).

The inorganic carbon concentration in ultramafic-influenced systems is highly dependent on pH and magmatic inputs. In low-pH ultramafic systems, concentrations of ΣCO2 can reach as high as those observed in magmatic systems, at ~4–20 mM (Table 16.1; see (Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119) for a review). The δ13C values of this CO2 display “typical” mid-ocean ridge values of –4‰ to –2‰ in some cases such as the Rainbow vent field (Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119). In other locations such as the Logatchev field, it is unusually positive, up to +9.5‰, even in fluids with ΣCO2 concentrations higher than seawater (Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119,Reference Konn, Charlou, Holm and Mousis180). In alkaline ultramafic systems such as Lost City, the high pHs lead to the rapid precipitation of calcium carbonate and therefore vanishingly low concentrations of ΣCO2 in end-member fluids (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112,Reference Kelley, Karson, Früh-Green, Yoerger, Shank and Butterfield140). This removal likely occurs throughout the fluid circulation pathway. Carbonate mineralization is common in ultramafic rocks (Reference Bonatti, Lawrence, Hamlyn and Breger181), and isotope signatures indicate precipitation occurs both at cold seawater temperatures and at warmer (65–95°C) temperatures, where δ13C values indicate that the source ΣCO2 has a substantial mantle component (Reference Eickmann, Bach, Rosner and Peckmann107).

Methane concentrations in ultramafic systems are frequently an order of magnitude higher than those in unsedimented, basalt-hosted systems (Figure 16.5 and Table 16.1), and substantial methane anomalies along the MAR have been attributed to exports from these systems (Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119,Reference Rona, Widenfalk and Bostrom171Reference Rona, Bougault, Charlou, Appriou, Nelsen and Trefry173). Estimates from mantle 3He exports suggest serpentinization of ultramafic rocks could account globally for about 75% of the methane flux from mid-ocean ridge systems (Reference Keir182). Isotopic signatures point to a nonbiological source for this methane (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112,Reference Charlou, Donval, Konn, Ondreas, Fouquet, Jean-Baptiste and Rona119,Reference McCollom and Seewald183), although in most systems more CH4 is present than would be expected in thermodynamic equilibrium with CO2 (for reviews, see (Reference McCollom, Lowell, Seewald, Metaxas and Perfit101,Reference McCollom and Seewald183)). One possibility is that the methane was formed long ago, at higher temperatures than the present day, and is subsequently stripped from vesicles in the rocks (Reference McDermott, Seewald, German and Sylva136,Reference Wang, Reeves, McDermott, Seewald and Ono184), which contain high CH4 and CO2 contents (Reference Kelley and Früh-Green185,Reference Kelley and Früh-Green186). Biologically derived methane from methanogenesis may also contribute (Reference Bradley and Summons187), albeit at relatively low levels when compared to the dominant nonbiological signature.

Short-chain hydrocarbons such as ethane, propane, and butane have been found in the low-micromolar concentrations in a wide range of ultramafic systems (Table 16.1) (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112,Reference McDermott, Seewald, German and Sylva136,Reference Konn, Charlou, Holm and Mousis180). Isotopic values that decrease with increasing chain length have been used to demonstrate that these species are not derived from the decomposition of sediments and could have a nonbiological origin such as Fischer–Tropsch-type reactions (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112). At the Lost City and Von Damm hydrothermal fields, the concentrations of these compounds increase in conjunction with methane concentrations (Reference Proskurowski, Lilley, Seewald, Früh-Green, Olson and Lupton112,Reference McDermott, Seewald, German and Sylva136), indicating similar processes may lead to their formation and/or cycling.

Formate and acetate have been reported in elevated concentrations in multiple ultramafic systems including Lost City (formate: 36–158 µM; acetate: 1–35 µM) (Reference Lang, Butterfield, Schulte, Kelley and Lilley141), Von Damm (formate: below detection to 669 µM) (Reference McDermott, Seewald, German and Sylva136), and Prony Bay (formate: ~4 µM; acetate: ~70 µM) (Reference Pisapia, Gerard, Gerard, Lecourt, Lang and Pelletier188). At Lost City, the isotopic composition of formate indicates it is synthesized by two pathways: abiotic synthesis in the subsurface that results in 14C-free formate with a δ13C signature (–13.0‰ to –8.9‰) similar to methane and short-chain hydrocarbons; and near-surface biological synthesis that incorporates modern DIC, resulting in formate with substantial 14C and a more positive δ13C signature (–9.1‰ to –4.3‰) (Reference Lang, Früh-Green, Bernasconi, Brazelton, Schrenk and McGonigle189). At the Von Damm vent field, higher concentrations of formate are found in hot mixed fluids than in pure end-member hydrothermal fluids, demonstrating that this species forms abiotically on timescales of hours to days (Reference McDermott, Seewald, German and Sylva136). At Lost City, the δ13C of acetate (–27‰ to –17‰) could be attributed to a mixture of anaerobic fermentation and acetogenesis (Reference Lang, Butterfield, Schulte, Kelley and Lilley141,Reference Lang, Früh-Green, Bernasconi, Brazelton, Schrenk and McGonigle189). Given the high abundances of microorganisms in the chimneys, the acetate could also be due to a thermocatalytic breakdown of complex organics in the biomass (Reference Lang, Butterfield, Schulte, Kelley and Lilley141,Reference Lang, Früh-Green, Bernasconi, Brazelton, Schrenk and McGonigle189).

Hydrolyzable amino acids are present in high abundances in Lost City fluids and chimneys (Reference Lang, Früh-Green, Bernasconi and Butterfield142). In the fluids, the highest concentrations were observed in locations where concentrations of H2 had been drawn down by sulfate reducers living in the sub-seafloor or chimney. The 13C of amino acids isolated from the chimneys had fractionation patterns consistent with synthesis by a chemolithoautotrophic source (Reference Lang, Früh-Green, Bernasconi and Butterfield142). In high-temperature fluids (>300°C) from Rainbow and Ashadze, dissolved free amino acids were detected in the picomolar concentration range, with tryptophan, phenylalanine, and leucine detected in the fluids but not in deep seawater (Reference Konn, Charlou, Holm and Mousis180). Tryptophan and phenylalanine contain aromatic rings that may assist in molecular stability at high temperatures (Reference Kawka and Simoneit190).

16.3.6 Fluxes between the Ocean and Crust

Hydrothermal circulation is the primary means of transferring materials between the crust and ocean (Reference Elderfield and Schultz78,Reference Kadko, Baross, Alt, Humphris, Zierenberg, Mullineaux and Thomson191). The net flux of constituents includes both input and removal processes, though these may be geographically and temporally distinct. The impact of hydrothermal circulation on the carbon budget of the ocean remains unconstrained in many ways.

Inorganic carbon is transferred to the deep ocean via magma degassing and removed by carbonate precipitation in the sub-seafloor at roughly similar rates (Figures 16.4 and 16.5). Degassing of mantle volatiles through high-temperature venting is estimated to input ~1 × 1012 mol C yr–1 of ΣCO2 into the ocean (Reference Kadko, Baross, Alt, Humphris, Zierenberg, Mullineaux and Thomson191,Reference Saal, Hauri, Langmuir and Perfit192). Carbonate precipitation is estimated to remove 1–3 × 1012 mol C yr–1 in ridge flanks (Reference Sansone, Mottl, Olson, Wheat and Lilley167,Reference Staudigel, Hart, Schmincke and Smith193,Reference Alt and Teagle194). Seawater passing through peridotites results in a loss of 0.4–2.0 × 1011 mol C yr–1, although stable isotope signatures indicate that approximately half of the carbon sequestered into the rock is in the form of organic carbon (Reference Schwarzenbach, Lang, Frah-Green, Lilley, Bemasconi and Mehay195,Reference Alt, Schwarzenbach, Frueh-Green, Shanks, Bernasconi and Garrido196).

Export and removal fluxes of DOC can be estimated by combining changes in concentrations with water fluxes through different types of hydrothermal systems (Reference Johnson and Pruis75,Reference Stein and Stein80,Reference Mottl, Halbach, Tunnicliffe and Hein163). If high-temperature vents remove an average of ~20 µmol of seawater DOC per liter, approximately 0.7–1.4 × 1010 g C yr–1 would be lost globally (Reference Lang, Butterfield, Lilley, Johnson and Hedges116). A similar scale loss of 1.4 ± 0.7 × 1010 g C yr–1 has been estimated based on changes of SPE-DOC concentrations (Reference Hawkes, Rossel, Stubbins, Butterfield, Connelly and Achterberg155). Ridge flank regions where crustal temperatures are “warm” (>45°C) have more substantial chemical changes in circulating fluids but smaller fluid fluxes than regions where crustal temperatures are “cool” (Reference Mottl and Wheat77,Reference Mottl, Halbach, Tunnicliffe and Hein163). If concentrations from the “warm” Juan de Fuca ridge flank system are typical of such systems, 2–13 × 1010 g C yr–1 would be removed (Reference Lang, Butterfield, Lilley, Johnson and Hedges116). Due to the larger water fluxes, if DOC concentrations through the “cool” crust at North Pond are globally representative, losses would be an order of magnitude higher at ~9–14 × 1011 g C yr–1 or ~5% of the total annual deep oceanic DOC loss (Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135).

16.4 Sedimented Hydrothermal Systems

Where spreading centers occur under thick sediment packages, hot water rapidly alters the OM fueling heterotrophic communities (e.g. (Reference Lin, Koch, Feseker, Ziervogel, Goldhammer and Schmidt125,Reference Wellsbury, Goodman, Barth, Cragg, Barnes and Parkes156,Reference Kawka and Simoneit190)), releasing inorganic carbon (Reference Seewald, Cruse and Saccocia89), influencing local physiochemical conditions, or forming complex oil-like materials (e.g. (Reference Simoneit and Lonsdale197,Reference Kvenvolden and Simoneit198)). The form and fate of carbon in heated sediments depend on its origin (terrigenous versus marine versus chemoautotrophic), temperature, and flow rate. Upon heating, a series of reactions similar to those that give rise to petroleum proceeds, with important differences due to the more water-rich conditions. The production of petroleum is generally considered to begin at ~50–70°C (Reference Tissot and Welte199). Weak bonds that sorb organic molecules onto surfaces break most easily, followed by bonds involving oxygen, sulfur, or nitrogen. Carbon–carbon bonds require the most energy – and therefore greater temperature or time – to break (Reference Tissot and Welte199).

Small polar compounds can be mobilized through enhanced desorption and the destruction of noncovalent bonds. The most labile material is removed from the solid phase due to microbial activity, pyrolysis, and/or desorption (Reference Wellsbury, Goodman, Barth, Cragg, Barnes and Parkes156). Over time, the amount of OM transferred into the aqueous phase decreases as the material is physically transported out of the system or biodegraded by microorganisms (Reference Lin, Koch, Feseker, Ziervogel, Goldhammer and Schmidt125,Reference Wellsbury, Goodman, Barth, Cragg, Barnes and Parkes156,Reference Seewald, Seyfried and Thornton157).

Unlike the dry “cracking” reactions that dominate petroleum reservoirs, breaking carbon–carbon bonds in the presence of water results in more oxidized products. Cracking reactions proceed at temperatures above ~100°C and result in CH4 and low-molecular-weight hydrocarbons (Reference Tissot and Welte199Reference Lewan, Winters and McDonald201). In contrast, in the presence of water and minerals, n-alkanes will instead degrade to oxygenated products such as alcohols, ketones, carboxylic acids, and, ultimately, CO2 and CH4 (Reference Seewald, Cruse and Saccocia89,Reference Seewald202). Sediments heated in aqueous environments produce copious amounts of acetate in particular. The reaction temperature impacts the products, with higher temperatures favoring more oxidized products such as CO2 over CH4 and propanol over propane (Reference Shock, Canovas, Yang, Boyer, Johnson and Robinson203).

Reduction, condensation, and dehydration reactions proceed at higher temperatures to form macromolecules and aromatics, causing compounds to revert to their most stable states (Figure 16.4) (Reference Simoneit204,Reference Didyk and Simoneit205). Polycyclic aromatic hydrocarbons and cyclic polysulfides, major components of some hydrothermal oils, form only under very high heat (>~300°C) and are signatures of elevated temperatures (Reference Kawka and Simoneit190,Reference Simoneit and Lonsdale197,Reference Kvenvolden and Simoneit198). Polypeptides form through dehydration and reduction, while lipids crack and recombine (Reference Rushdi and Simoneit206).

Water washing will selectively transport more soluble components from the subsurface to the surface and leave behind larger condensates (Reference Lin, Koch, Feseker, Ziervogel, Goldhammer and Schmidt125,Reference Kawka and Simoneit190,Reference Tissot and Welte199,Reference Simoneit, Lein, Peresypkin and Osipov207). Smaller alkanes (<C10), aromatic volatiles, compounds containing C–N–S bonds, oligosaccharides, and oligopeptides are often missing in sediments subjected to “water washing,” while fluids and plumes contain higher concentrations of these compounds (Reference Kawka and Simoneit190,Reference Simoneit204,Reference Simoneit, Lein, Peresypkin and Osipov207).

These released compounds are highly biodegradable and fuel heterotrophic organisms. The labile amino acids released from sterilized sediments, for example, are utilized and reworked by microorganisms in parallel, nonsterilized experiments (Reference Lin, Koch, Feseker, Ziervogel, Goldhammer and Schmidt125). In general, the low-molecular-weight organic acids that are primary breakdown products of heating sediments in the presence of water, particularly acetate, are important substrates for anaerobic microorganisms (Reference Parkes, Taylor and Jorckramberg208).

The residual OM that is not removed with water washing is enriched in less soluble material, leading to “hydrothermal petroleum.” Cooling near the sediment–water interface can help trap less soluble compounds through differential condensation and solidification (Reference Kawka and Simoneit190,Reference Simoneit, Lein, Peresypkin and Osipov207,Reference Simoneit, Kawka and Brault209). The distribution of compounds and the maturity of these oils are highly variable.

16.5 Continental Subsurface

Geological heterogeneity produced through plate tectonics diversifies and segments the continental deep subsurface and its constituent biospheres differently from in the marine realm. Mountain and basin formation juxtaposes reactive rocks and minerals and creates new hydrological flow paths. Rock and water ages on the continents range from modern to billions of years (Reference Murdoch, Germanovich, Wang, Onstott, Elsworth and Stetler210,Reference Holland, Lollar, Li, Lacrampe-Couloume, Slater and Ballentine211). Terrestrial vegetation supplies vast quantities of organic carbon, although this influence is attenuated with increasing depth. The water age, hydrological connectivity, and major element chemistry of continental subsurface sites dramatically impact carbon cycling and the nature of in situ biospheres.

The continental deep subsurface extends downward from the base of the critical zone (Reference Pedersen212,Reference Colwell and D’Hondt213), although specific depths and thresholds have yet to be defined, particularly on the upper boundary. The penetration of life into the continental crust appears to be limited not strictly by depth, but rather by temperature, permeability, and perhaps aridity, with clear life detection in even the deepest boreholes and mines. Sites lacking identifiable life are few and far between and appear to be limited by temperature (e.g. German continental deep drilling program (KTB) cores in the Black Forest (Reference Huber, Huber, Lüdemann and Stetter214)) or aridity (Reference Colwell and D’Hondt213).

Estimates of the size of the continental deep biosphere are large (ranging from 2.3 × 1015 to 1017 g C), mirroring similar estimates of the marine deep biosphere (4.1 × 1015 g C) and rivaling terrestrial soils (2.6 × 1016 g C) (Reference Whitman, Coleman and Wiebe215Reference McMahon and Parnell217). The uncertainty in these calculations spans orders of magnitude and has not changed significantly since the original estimates by Whitman (Reference Whitman, Coleman and Wiebe215; Chapter 17, this volume), although the trend is downward (Chapter 17, this volume). However, increasing levels of inquiry applied globally using advanced methodologies have identified abundant, taxonomically diverse communities within the continental deep subsurface, giving credence to vast amounts of carbon contained and being cycled by these ecosystems (Reference Fredrickson and Onstott218Reference Probst, Castelle, Singh, Brown, Anantharaman, Sharon, Hug, Burstein, Emerson, Thomas and Banfield222).

16.5.1 Types of Continental Deep Subsurface Environments

Continental deep subsurface environments can be broadly divided between sedimentary and crystalline host rocks, but even within this framework they range significantly in carbon content, isolation from the surface, and dominant carbon cycling processes (Figure 16.1). The best-studied sites are found in shallow sedimentary and igneous aquifers owing largely to their relevance to human water supplies (Reference Stevens and McKinley223Reference Murphy, Schramke, Fredrickson, Bledsoe, Francis and Sklarew231). Hydrocarbon reservoirs contain vast quantities of organic carbon and have distinct microbiology associated with their formation waters (Reference Means and Hubbard232Reference Aitken, Jones and Larter234). A recent emphasis on deep coal beds and their constituent carbon cycling has come to the scientific forefront due to their importance in gas extraction via deep fracking technologies (Reference Strapoc, Mastalerz, Dawson, Macalady, Callaghan and Wawrik235). Deep crystalline bedrock sites feature the oldest, deepest, and most isolated deep biosphere environments (Reference Holland, Lollar, Li, Lacrampe-Couloume, Slater and Ballentine211,Reference Pedersen, Fredrickson and Fletcher236Reference Lollar, Onstott, Lacrampe-Couloume and Ballentine238). Caves, in contrast, sit at the interface between the surface and the deep and are covered more completely in other reviews (Reference Boston, Spilde, Northup, Melim, Soroka and Kleina239Reference Barton and Northup241). This section will describe the forms, cycling, and fate of organic carbon in each environment.

While the deep subsurface biosphere is pervasive, it is difficult to access reliably. Common access points are wells, boreholes, mines, and caves. Each approach has the potential to impact in situ processes and must be considered when evaluating data sets. Natural springs are often considered as “portals” or “windows” into the deep biosphere, often showing a mix of surface and subsurface communities (Reference Suzuki, Ishii, Wu, Cheung, Tenney and Wanger242,Reference Magnabosco, Tekere, Lau, Linage, Kuloyo and Erasmus243). The last 10 years has seen the establishment of a number of deep subsurface observatories including into permafrost (Permafrost Tunnel Research Facility, AL, USA), deep crystalline bedrock (Deep Mine Microbial Observatory (DeMMO), SD, USA; Coast Range Ophiolite Microbial Observatory (CrOMO), CA, USA; Äspö Hard Rock Laboratory, Sweden; and many others), and sedimentary aquifers (Deep Biosphere in Terrestrial Systems (DEBITS), New Zealand; Savannah River Site, SC, USA).

16.5.2 Continental Carbon Cycling

Organic carbon in the continental deep biosphere may derive from surficial inputs, in situ autotrophic carbon fixation, water–rock reactions, or ancient sedimentary sources. The relative balance of these sources depends sharply on geology, both by surface connectivity and host lithology. The following sections describe this balance in sedimentary and igneous aquifers, hydrocarbon reservoirs, deep coal beds, and deep crystalline bedrock.

Key processes in subsurface carbon cycling depend on the relative recalcitrance of ancient OM, supply of labile organic carbon, input of metabolic oxidants and reductants, and aquifer porosity and permeability. Microbial carbon fixation produces labile organic carbon and methane, whereas heterotrophic microbial processes consume both labile and recalcitrant subsurface organic carbon. The relative importance of these two end members broadly suggests autotrophic processes dominate in crystalline and deep rock aquifers whereas heterotrophic processes are more abundant within sedimentary systems, although numerous counterexamples exist and both processes (e.g. (Reference Stevens251)) must be active for a functioning ecosystem (Reference Fredrickson and Balkwill225). Organic acids and short-chain hydrocarbons are key microbial products and substrates within most continental subsurface settings with typical concentrations in the 10–100 µM range (Table 16.2). Due to rock dissolution and other processes, DIC can be very high, and may get much higher as aquifers are targeted for anthropogenic carbon sequestration (Reference Probst, Castelle, Singh, Brown, Anantharaman, Sharon, Hug, Burstein, Emerson, Thomas and Banfield222).

Table 16.2 Summary of characteristics and carbon contents of different types of continental subsurface systems.

System typeShallow sedimentary aquifersShallow igneous aquifersHydrocarbon reservoir formation watersCoal bedsDeep bedrock
Example systemsLower Saxony, GermanyaColumbia River flood basaltsbPalo Duro BasincGerman lignite depositsdSouth African gold minese
Defining characteristicsShallow confined and unconfined aquifers with abundant sedimentary OM, fresh watersThick basalt deposits with confined aquifers, interbedded sediments, fracture-based porosity, and relatively fresh NaCl-dominated fluidsSedimentary hydrocarbons interfacing with aqueous brines, organic alkalinity may exceed bicarbonate alkalinityExtremely organic-rich sediments at varying stages of thermal maturity, limited porosity and permeabilityDeep (1.0–3.3 km) fracture-based fluids with thousand- to million-year recharge times
TemperaturesLow to moderateLow to moderateModerate to highLow to moderateLow to high
Recharge timescaleRapid to moderateRapid to moderateModerate to longModerate to longModerate to very long
pHCircumneutral7.5–8.5 shallow,
8.0–10.5 deep
5–86.8–7.27.4–9.4
H2 (mM)Up to 0.067.4
SO4 (mM)Generally <0.5, but up to 2Up to 25Up to 0.1480.623
Total DIC (mM)0.125–2.80019.8–43.60.09–2.40
δ13CCO2 (‰)–20–30 to 20; mostly –10–14 to 20–43 to –5
CH4 (mM)0.00089–2.68000Up to 160Very high0.010–0.1000.026–8.800
δ13CCH4–110 to 20 (mean –70)Variable–81 to –71–58 to –37
Short-chain hydrocarbons Σ(C2H6–C4H8) (µM)~3 (median)High<0.1–201
DOC (mM)0.17–0.300.16–0.390.05–14.750.19–0.95n.d. to 0.410
Formate2.22–31.1 µmol/g sed0.44–34.00 µM
Acetate6.08 mM1.7–8.5 µmol/g sedf0.07–28.00 µM
Amino acids0.0133–2.6700 µM
Typical cell density (cells/mL)103–108103–105High10710–105

16.5.3 Sedimentary and Igneous Aquifers

Both sedimentary (e.g. Atlantic coastal plain) and igneous aquifers (e.g. Columbia River basalt aquifer) have been shown to contain vibrant microbial communities and have been the subject of intensive study due to their economic and social importance as sources of drinking and industrial water, as well as their vulnerability to anthropogenic contamination (Reference Fredrickson and Balkwill225,Reference Lapworth, Gooddy, Butcher and Morris226,Reference Murphy, Schramke, Fredrickson, Bledsoe, Francis and Sklarew231). Recharge timescales of aquifers vary over many orders of magnitude (months to millions of years), controlling the relative supply of exogenous DOC and electron acceptors. In many systems, significant supply of young sedimentary carbon produces relatively high DOC, methane, and organic acid concentrations. The composition of this DOC can be complex, including significant amounts of nitrogen- and sulfur-bearing organic molecules (Reference Longnecker and Kujawinski252). While oligotrophic compared to surface environments, aquifers are relatively carbon rich for the subsurface and can support correspondingly high cell densities (e.g. 105 cells/mL), even in oligotrophic crystalline aquifers (Reference Fry, Fredrickson, Fishbain, Wagner and Stahl246).

Primary productivity within aquifers varies tremendously based on exogenous and sedimentary organic carbon supply, but is significant in some settings. Hydrogen production can be large and may support autotrophic populations, particularly in igneous and ultrabasic host environments, fueling the so-called subsurface lithoautotrophic microbial ecosystems (Reference Stevens and McKinley223,Reference Stevens251,Reference Chapelle, O’Neill, Bradley, Methe, Ciufo and Knobel253). Utilization of iron oxide minerals as terminal electron acceptors for both autotrophic and heterotrophic metabolisms is common, producing high concentrations of dissolved ferrous iron in many groundwaters (Reference Griebler and Lueders254,Reference Flynn, Sanford, Ryu, Bethke, Levine and Ashbolt255). Iron and sulfur oxidative metabolisms are also found where microaerophilic conditions or sufficient nitrate concentrations exist (Reference Griebler and Lueders254,Reference Probst, Ladd, Jarett, Geller-McGrath, Sieber and Emerson256). Sulfate is a less dominant anion in continental settings relative to its ubiquity in the marine realm, but where present, it can fuel significant populations of autotrophic and heterotrophic sulfate-reducing bacteria (Reference Griebler and Lueders254,Reference Flynn, Sanford, Ryu, Bethke, Levine and Ashbolt255).

Heterotrophic processes rely on the input of DOC from the mineralization of sedimentary carbon, aquifer recharge, or in situ microbial activity. In 85% of US aquifers, DOC concentrations were <175 µM (median 42 and 58 µM). These ranges were not significantly different between sedimentary and crystalline aquifers (ranging from 8 to 275 µM, median 42 µM) (Reference Leenheer, Malcolm and McKinley257). A more recent analysis of DOC concentration across UK aquifers showed a range of 15–1550 µM (257 µM average) (Reference Lapworth, Gooddy, Butcher and Morris226), although this sampling includes evidence for significant contamination from agriculture and concomitant microbial respiration that introduced OM. Locally, high concentrations of organic acids (up to 60 µM formate) can be produced by microbial degradation of complex sedimentary OM, particular in shale horizons, which may then diffuse to more porous sediments, driving respiration (Reference McMahon and Chapelle258). For shallow aquifer systems, periodic environmental changes related to seasonal shifts, water table fluctuations, or land use may transport both DOC and oxidants to depth, driving increases in heterotrophic respiration (Reference Baker, Valett and Dahm244,Reference Kusel, Totsche, Trumbore, Lehmann, Steinhauser and Herrmann259).

Methane is a ubiquitous reservoir of organic carbon in sedimentary aquifers. Methane concentrations are extremely variable but sometimes can reach extremely high values (e.g. 0.9 nM to 2.7 mM in the Lower Saxony region of Germany (Reference Schloemer, Elbracht, Blumenberg and Illing229) and 3.1 nM to 293 µM across Great Britain (Reference Bell, Darling, Ward, Basava-Reddi, Halwa and Manamsa230)). Sources of methane vary and include abiotic and biotic sources, including microbial methanogenesis (including hydrogenotrophic, acetoclastic, and methyl fermentation) as well as thermogenic cracking of buried OM (Reference Veto, Futo, Horvath and Szanto224,Reference Schloemer, Elbracht, Blumenberg and Illing229,Reference Bell, Darling, Ward, Basava-Reddi, Halwa and Manamsa230). The isotopic composition of methane and co-occurring short-chain hydrocarbons can be used to assess methane sources and suggest active microbial CO2 reduction as the primary source in both German and British aquifers (Reference Schloemer, Elbracht, Blumenberg and Illing229,Reference Bell, Darling, Ward, Basava-Reddi, Halwa and Manamsa230). High concentrations tend to correlate to organic-rich, low-SO4 geological formations (Reference Schloemer, Elbracht, Blumenberg and Illing229).

16.5.4 Hydrocarbon Reservoirs

Hydrocarbon reservoirs were among the earliest studied continental deep biospheres, with experiments beginning in the 1920s by Colwell and D’Hondt (Reference Colwell and D’Hondt213). These systems are characterized by large accumulations of liquid and gaseous hydrocarbons, providing abundant sources of carbon and electron donors, but they tend to be correspondingly depleted in oxidants and nutrients. Extremely high concentrations of volatile organic acids (particularly acetate) comprise the majority of DOC in the water phases of hydrocarbon-bearing basins, reaching concentrations of up to hundreds of mM (Reference Means and Hubbard232,Reference Head, Jones and Larter260,Reference Nazina, Shestakova, Ivoilov, Kostrukov, Belyaev and Ivanov261).

The most significant metabolisms in hydrocarbon reservoirs are sulfate reduction, methanogenesis, acetogenesis, iron reduction, and fermentation (Reference Head, Jones and Larter260Reference Magot, Ollivier and Patel262), the balance of which is determined by electron acceptor supply. Spatially, biodegradation of oil is concentrated at the oil–water interface and is limited by reservoir temperature, with limited activity being observed above 80°C (Reference Aitken, Jones and Larter234,Reference Larter, Huan, Adams, Bennett, Jokanola and Oldenburg263). Anthropogenic influence through drilling, water introduction, casing, and fracturing of reservoirs and the introduction of exogenous microbes can significantly change in situ carbon cycling, most notoriously causing reservoir souring by stimulating sulfate-reducing populations in previously methanogenic reservoirs. For more a complete description of carbon cycling and biodegradation in hydrocarbon reservoirs, see the reviews by Larter et al. (Reference Larter, Huan, Adams, Bennett, Jokanola and Oldenburg263), Means and Hubbard (Reference Means and Hubbard232), and Head et al. (Reference Head, Jones and Larter260).

16.5.5 Deep Coal Beds

Coal is formed through the burial and diagenesis of large accumulations of terrestrial plant matter and therefore contains an extremely high organic carbon content. The bioavailablity of this OM to deep subsurface microbes depends on thermal maturity and burial history, which control the form and speciation of OM as well as the sterilization history of resident microbial populations (Reference Strapoc, Mastalerz, Dawson, Macalady, Callaghan and Wawrik235). Low-maturity (rank) coals are the most bioavailable and actively accumulate biogenic methane. Aqueous extracts of low-maturity coals and lignites produce extremely high concentrations of organic acids, including acetate, formate, and oxalate in the range of 0.37–2.5.0 mg/g sediment (Reference Vieth, Mangelsdorf, Sykes and Horsfield247,Reference Zhu, Vieth-Hillebrand, Wilke and Horsfield264). Yields of labile OM decrease significantly with increasing thermal maturity (Reference Vieth, Mangelsdorf, Sykes and Horsfield247,Reference Zhu, Vieth-Hillebrand, Wilke and Horsfield264).

Biogenic methane appears to universally accumulate in coals at <80°C (Reference Strapoc, Mastalerz, Dawson, Macalady, Callaghan and Wawrik235). Microbial processing of coal to methane is a multistep process that requires and supports an ecosystem of microbes. First, organic polymers are fragmented into hydrocarbon intermediates, followed by a secondary fermentation to methanogenic substrates like CO2, H2, organic acids, and alcohols. These substrates then fuel acetoclastic, methylotrophic, and hydrogenotrophic methanogens (Reference Strapoc, Mastalerz, Dawson, Macalady, Callaghan and Wawrik235). The rate and efficiency of these processes in different coal deposits and the accessibility of this methane for extraction are of considerable economic importance. For a more complete review of coal bed biogeochemistry, see Strapoc et al. (Reference Strapoc, Mastalerz, Dawson, Macalady, Callaghan and Wawrik235).

16.5.6 Deep Bedrock

Deep crystalline bedrock-hosted biospheres stand in contrast to the aforementioned settings in their constituent reservoirs and fluxes of carbon and energy. Here, inputs from the surface are limited, with water residence times reaching millions to billions of years (e.g. (Reference Holland, Lollar, Li, Lacrampe-Couloume, Slater and Ballentine211,Reference Onstott, Lin, Davidson, Mislowack, Borcsik and Hall265)) and sedimentary carbon (where present) is recalcitrant to graphitic carbon (Reference Kietavainen, Ahonen, Niinikoski, Nykanen and Kukkonen266). The largest pool of organic carbon is often as methane, although considerable variability is present with depth and lithology (Reference Simkus, Slater, Lollar, Wilkie, Kieft and Magnabosco248,Reference Kietavainen, Ahonen, Niinikoski, Nykanen and Kukkonen266). Porosity and permeability is fracture based, adding a stochastic temporal dynamic to fluxes and mixing (Reference Onstott, Lin, Davidson, Mislowack, Borcsik and Hall265,Reference Lollar, Voglesonger, Lin, Lacrampe-Couloume, Telling and Abrajano267).

Hydrogen, methane, sulfate, and iron cycling drive primary production in deep crystalline bedrock settings. The relative importance of these processes is variable with depth, host lithology, and fluid chemistry (Reference Osburn, LaRowe, Momper and Amend221,Reference Simkus, Slater, Lollar, Wilkie, Kieft and Magnabosco248). Precambrian rocks, which constitute the best-studied deep crystalline biospheres, are prolific producers of hydrogen (up to mM concentrations) (Reference Lollar, Onstott, Lacrampe-Couloume and Ballentine238,Reference Lollar, Voglesonger, Lin, Lacrampe-Couloume, Telling and Abrajano267), which can serve as the terminal electron donor for either sulfate reduction or CO2 reduction-based metabolisms (Reference Lau, Kieft, Kuloyo, Linage-Alvarez, Van Heerden and Lindsay249,Reference Onstott, Lin, Davidson, Mislowack, Borcsik and Hall265,Reference Suzuki, Konno, Fukuda, Komatsu, Hirota and Watanabe268Reference Magnabosco, Ryan, Lau, Kuloyo, Lollar and Kieft272).

Analysis of subsurface genomes shows that enzymes of hydrogen metabolism are overrepresented, emphasizing the potential dominance of this metabolic strategy (Reference Colman, Poudel, Stamps, Boyd and Spear273). Metagenomic surveys suggest that carbon fixation is performed primarily using the reductive acetyl CoA pathway (Reference Probst, Ladd, Jarett, Geller-McGrath, Sieber and Emerson256,Reference Momper, Jungbluth, Lee and Amend271,Reference Magnabosco, Ryan, Lau, Kuloyo, Lollar and Kieft272). Extreme metabolic flexibility has been observed in the cosmopolitan subsurface dweller Candidatus Desulforudis audaxviator, which can grow in near monoculture in isolated fracture systems (Reference Chivian, Brodie, Alm, Culley, Dehal and DeSantis220) and has been found globally (Reference Osburn, LaRowe, Momper and Amend221,Reference Jungbluth, del Rio, Tringe, Stepanauskas and Rappe274).

Heterotrophic microbes and metabolisms have been found to dominate in some deep crystalline settings, despite the apparent limited availability of exogenous carbon. Abiogenic sources of methane in addition to limited populations of microbial methanogens supply a significant flux of methane to fuel methanotrophic communities reaching tens of mM concentrations (Reference Simkus, Slater, Lollar, Wilkie, Kieft and Magnabosco248,Reference Kietavainen, Ahonen, Niinikoski, Nykanen and Kukkonen266,Reference Lollar, Lacrampe-Couloume, Slater, Ward, Moser and Gihring275,Reference Kadnikov, Frank, Mardanov, Beletskii, Ivasenko and Pimenov276). Methane cycling has been observed to be most active at moderate depths (0.5–1.5 km) rather than in the deepest, most isolated settings (Reference Simkus, Slater, Lollar, Wilkie, Kieft and Magnabosco248,Reference Kietavainen, Ahonen, Niinikoski, Nykanen and Kukkonen266,Reference Momper, Reese, Zinke, Wanger, Osburn and Moser277). Other sources of carbon for heterotrophic communities include biofilm-based small organic compounds (Reference Onstott, Lin, Davidson, Mislowack, Borcsik and Hall265,Reference Wanger, Southam and Onstott278), free organic acids formed through fermentation or abiogenesis, and ancient organic carbon (Reference Purkamo, Bomberg, Nyyssonen, Kukkonen, Ahonen and Itavaara279).

Mineral and biofilm-based metabolisms may be particularly important in deep crystalline settings. Increasing evidence of extensive adaptation to life in biofilms is emerging from these environments in the form of physical adaptations like grappling appendages observed in putative CandidatusAltiarchaeum” (Reference Probst, Ladd, Jarett, Geller-McGrath, Sieber and Emerson256), as well extracellular electron transport in subsurface isolates (Reference Jangir, French, Momper, Moser, Amend and El-Naggar280). In high-pH settings, autotrophic populations may depend on solid carbonate minerals due to carbon speciation in ultrabasic environments (Reference Suzuki, Kuenen, Schipper, van der Velde, Ishii and Wu269). Differences between the attached and planktonic communities have long been observed in crystalline aquifer settings (Reference Osburn, LaRowe, Momper and Amend221,Reference Momper, Reese, Zinke, Wanger, Osburn and Moser277,Reference Lehman, Colwell and Bala281), often with orders of magnitude higher cell densities present within the biofilms (Reference Wanger, Southam and Onstott278). The net suggestion of these observations is that mineral and biofilm-based lifestyles are the norm for the deep continental subsurface, but are as yet undersampled. Efforts to cultivate and characterize the metabolic capacities of these attached communities are underway.

16.6 Conclusion

16.6.1 Broad Similarities across Systems

The deep biosphere spans an incredible range of physical and chemical conditions. Despite their heterogeneity, some broad similarities are present across systems. Organic carbon concentrations reach their highest levels in regions that have large inputs from primary producers, either presently (continental margins, shallow sedimentary aquifers, diffuse hydrothermal vents) or in the past (hydrocarbon reservoirs, coal beds). In contrast, concentrations are lower in rocky areas with little sedimentary input and low amounts of chemolithoautotrophy in both continental (shallow igneous aquifers, deep bedrock) and marine (ridge flanks, high-temperature hydrothermal vents) systems. Elevated concentrations of methane are related to the anaerobic breakdown of OM and associated methanogenesis (sedimentary porewaters, sedimented hydrothermal systems, hydrocarbon reservoirs, coal beds), but also due to hydrogenotrophic methanogenesis, with the hydrogen supplied by water–rock reactions (basalts) and from mantle inputs (hydrothermal systems). Somewhat surprising is the persistence of some forms of organic carbon that are generally thought to be readily accessible to microorganisms, such as acetate, in many systems (sedimentary porewaters, shallow igneous aquifers, deep bedrock).

16.6.2 Limits to Knowledge and Unknowns

(1) Exchange/transformation of carbon between aqueous and solid phases. A characteristic of subsurface environments is the ubiquitous presence of a solid phase, be it from surface-derived particles or crystalline rocks. The exchange and transformation of carbon between the aqueous and solid phases is therefore a major mechanism for controlling the form and fate of carbon in the subsurface. Major questions remain as to what controls these exchanges and the degree to which they are catalyzed by minerals and microorganisms.

(2) Bioavailability of organic carbon. Microbial respiration has been invoked to account for the oxidation of OM that is millions of years old in sediments (Reference Røy67) and the removal of oceanic dissolved OM that is thousands of years old in the basaltic basement (Reference Walter, Jaekel, Osterholz, Fisher, Huber and Pearson135). Even 365 million-year-old shale carbon can be incorporated into cellular biomass given the right conditions (Reference Petsch, Eglinton and Edwards282). These studies raise the intriguing question of whether all OM is, ultimately, bioavailable give enough time and a favorable setting, or whether there is some pool that will resist remineralization to CO2 under all circumstances. This question ties directly into point (3) below.

(3) Controls on reaction rates of biogeochemical processes. The rate at which carbon is transformed or remineralized is fundamentally important to understanding the short- and long-term controls on the global carbon cycle and to identifying the distribution of subsurface life. The processes occurring in the marine and continental subsurface are inherently difficult to accurately mimic in laboratory experiments. While short-term experiments can address the more reactive portions of the organic pool, our understanding of the transformations that occur over century or millennium timescales, particularly when uncultured microorganisms mediate the reactions, is more challenging but no less important.

(4) Predictive ability. This review describes what types of carbon are present in distinct geological, geochemical, and biological environments. Ultimately, however, the reverse is a major goal: the ability to have such a fundamental grasp of the mechanistic controls on carbon cycling that it is possible to accurately predict what types and abundances of carbon will be present in a given system.

(5) Characterization of OM. Despite decades of effort and major progress on several fronts, the molecular structure of the vast majority of OM in the subsurface remains uncharacterized. This gap in our knowledge will continue to inhibit our understanding of carbon biogeochemical cycling in the subsurface.

Acknowledgments

Funding support for SQL is from C-DEBI (NSF grant OIA-0939564) and National Science Foundation grant OCE-1536702. Funding support for ADS is from C-DEBI and NSF OCE-1357242. Funding support for MRO is from NASA Exobiology (NNH14ZDA001N) and NAI Life Underground (NNH12ZDA002C). This is C-DEBI contribution 475.

Questions for the Classroom

1 Continental and oceanic subsurface crystalline aquifers are similar in many ways. How do the characteristics of carbon in, for example, the oceanic North Pond and the continental South African gold mine systems compare and differ? Why?

2 Are the sites studied thus far representative of globally relevant locations where carbon is processed in the subsurface? What locations or geological systems are missing? Why have these not yet been studied? What are the prospects for studying these locations?

3 What effect are humans having on carbon in the deep biosphere?

4 What are the next steps to improve our ability to computationally model different forms of carbon in the subsurface and how they change in the subsurface?

5 Imagine a hypothetical microorganism that is capable of remineralizing any type of nonbioavailable OM back to CO2. If this microorganism proliferated in subsurface environments, what would the effect be?

6 Ultraviolet radiation can create radical species (compounds that are highly reactive due to the presence of an unpaired electron) that can oxidize organic molecules via random reactions. What is the likely effect of a change in atmospheric ozone concentrations, and therefore ultraviolet flux to Earth’s surface, on organic carbon burial rates?

References

Martiny, JBH, Bohannan, BJM, Brown, JH, Colwell, RK, Fuhrman, JA, Green, JL, et al. Microbial biogeography: putting microorganisms on the map. Nature Reviews Microbiology. 2006;4(2):102–112.Google Scholar
Hedges, JI. Global biogeochemical cycles – progress and problems. Marine Chemistry. 1992;39(1–3):6793.CrossRefGoogle Scholar
Schrenk, M, Huber, J, Edwards, K. Microbial provinces in the subseafloor. Annual Review of Marine Science. 2010;2:279304.Google Scholar
Rex, MA, Etter, RJ, Morris, JS, Crouse, J, McClain, CR, Johnson, NA, et al. Global bathymetric patterns of standing stock and body size in the deep-sea benthos. Marine Ecology Progress Series. 2006;317:18.CrossRefGoogle Scholar
LaRowe, DE, Burwicz, E, Arndt, S, Dale, AW, Amend, JP. Temperature and volume of global marine sediments. Geology. 2017;45(3):275–278.CrossRefGoogle Scholar
Burdige, DJ. Burial of terrestrial organic matter in marine sediments: a re-assessment. Global Biogeochemical Cycles. 2005;19(4):GB4011.Google Scholar
Burdige, DJ. Preservation of organic matter in marine sediments: controls, mechanisms, and an imbalance in sediment organic carbon budgets? Chemical Reviews. 2007;107(2):467–485.CrossRefGoogle Scholar
Hedges, JI, Keil, RG. Sedimentary organic matter preservation – an assessment and speculative synthesis. Marine Chemistry. 1995;49(2–3):81115.Google Scholar
Zonneveld, KAF, Versteegh, GJM, Kasten, S, Eglinton, TI, Emeis, KC, Huguet, C, et al. Selective preservation of organic matter in marine environments; processes and impact on the sedimentary record. Biogeosciences. 2010;7(2):483511.Google Scholar
Blair, NE, Aller, RC. The fate of terrestrial organic carbon in the marine environment. Annual Review of Marine Science. 2012;4:401–423.Google Scholar
Keil, RG, Mayer, LM. Mineral matrices and organic matter. In: Holland, HD, Turekian, KK, eds. Treatise on Geochemistry, 2nd edn. Oxford: Elsevier, 2014, pp. 337–359.Google Scholar
Keil, R, Annual, R. Anthropogenic forcing of carbonate and organic carbon preservation in marine sediments. Annual Review of Marine Sciences, 2017;9:151–172.Google Scholar
Middelburg, JJ. Reviews and syntheses: to the bottom of carbon processing at the seafloor. Biogeosciences. 2018;15(2):413–427.Google Scholar
Hedges, JI, Keil, RG, Benner, R. What happens to terrestrial organic matter in the ocean? Organic Geochemistry. 1997;27(5–6):195212.Google Scholar
Blair, NE, Leithold, EL, Ford, ST, Peeler, KA, Holmes, JC, Perkey, DW. The persistence of memory: the fate of ancient sedimentary organic carbon in a modern sedimentary system. Geochimica et Cosmochimica Acta. 2003;67(1):6373.Google Scholar
Galy, V, Peucker-Ehrenbrink, B, Eglinton, T. Global carbon export from the terrestrial biosphere controlled by erosion. Nature. 2015;521(7551):204–207.CrossRefGoogle ScholarPubMed
Hilton, RG, Galy, A, Hovius, N, Horng, MJ, Chen, H. Efficient transport of fossil organic carbon to the ocean by steep mountain rivers: an orogenic carbon sequestration mechanism. Geology. 2011;39(1):71–74.Google Scholar
Goni, MA, Hatten, JA, Wheatcroft, RA, Borgeld, JC. Particulate organic matter export by two contrasting small mountainous rivers from the Pacific Northwest, USA. Journal of Geophysical Research – Biogeosciences. 2013;118(1):112–134.CrossRefGoogle Scholar
Bao, HY, Lee, TY, Huang, JC, Feng, XJ, Dai, MH, Kao, SJ. Importance of Oceanian small mountainous rivers (SMRs) in global land-to-ocean output of lignin and modern biospheric carbon. Scientific Reports. 2015;5:16217.Google Scholar
Opsahl, S, Benner, R, Amon, RMW. Major flux of terrigenous dissolved organic matter through the Arctic Ocean. Limnology and Oceanography. 1999;44(8):2017–2023.CrossRefGoogle Scholar
Dittmar, T, Kattner, G. The biogeochemistry of the river and shelf ecosystem of the Arctic Ocean: a review. Marine Chemistry. 2003;83(3–4):103–120.CrossRefGoogle Scholar
Raymond, PA, McClelland, JW, Holmes, RM, Zhulidov, AV, Mull, K, Peterson, BJ, et al. Flux and age of dissolved organic carbon exported to the Arctic Ocean: a carbon isotopic study of the five largest arctic rivers. Global Biogeochemical Cycles. 2007;21(4):GB4011.Google Scholar
Feng, XJ, Vonk, JE, van Dongen, BE, Gustafsson, O, Semiletov, IP, Dudarev, OV, et al. Differential mobilization of terrestrial carbon pools in Eurasian Arctic river basins. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(35):14168–14173.Google Scholar
Blair, NE, Carter, WD, Boehme, SE. Diagenetic isotope effects in an anoxic marine sediment. Abstracts of Papers of the American Chemical Society. 1991;201:34-GEOC.Google Scholar
Schmitz, RA, Daniel, R, Deppenmeier, U, Gottschalk, G. The Anaerobic Way of Life. Prokaryotes: A Handbook on the Biology of Bacteria, Vol. 2, 3rd edn: Ecophysiology and Biochemistry. Washington, DC: American Chemical Society, 2006, pp. 86101.Google Scholar
Blair, NE, Martens, CS, Desmarais, DJ. Natural abundances of carbon isotopes in acetate from a coastal marine sediment. Science. 1987;236(4797):66–68.Google Scholar
Gelwicks, JT, Risatti, JB, Hayes, JM. Carbon isotope effects associated with aceticlastic methanogenesis. Applied and Environmental Microbiology. 1994;60(2):467–472.Google Scholar
Hinrichs, KU, Hayes, JM, Bach, W, Spivack, AJ, Hmelo, LR, Holm, NG, et al. Biological formation of ethane and propane in the deep marine subsurface. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(40):14684–14689.Google Scholar
Heuer, VB, Pohlman, JW, Torres, ME, Elvert, M, Hinrichs, KU. The stable carbon isotope biogeochemistry of acetate and other dissolved carbon species in deep subseafloor sediments at the northern Cascadia Margin. Geochimica et Cosmochimica Acta. 2009;73(11):3323–3336.Google Scholar
Suess, E. Particulate organic–carbon flux in the oceans – surface productivity and oxygen utilization. Nature. 1980;288(5788):260–263.CrossRefGoogle Scholar
Abdulla, H, Burdige, D, Komada, T. Accumulation of deaminated peptides in anoxic sediments of Santa Barbara Basin. Geochimica et Cosmochimica Acta. 2018;223:245–258.Google Scholar
Jiao, N, Herndl, G, Hansell, D, Benner, R, Kattner, G, Wilhelm, S, et al. Microbial production of recalcitrant dissolved organic matter: long-term carbon storage in the global ocean. Nature Reviews Microbiology. 2010;8(8):593–599.Google Scholar
Lechtenfeld, OJ, Kattner, G, Flerus, R, McCallister, SL, Schmitt-Kopplin, P, Koch, BP. Molecular transformation and degradation of refractory dissolved organic matter in the Atlantic and Southern Ocean. Geochimica et Cosmochimica Acta. 2014;126:321–337.CrossRefGoogle Scholar
Hartnett, HE, Keil, RG, Hedges, JI, Devol, AH. Influence of oxygen exposure time on organic carbon preservation in continental margin sediments. Nature. 1998;391(6667):572–574.Google Scholar
Hee, CA, Pease, TK, Alperin, MJ, Martens, CS. Dissolved organic carbon production and consumption in anoxic marine sediments: a pulsed-tracer experiment. Limnology and Oceanography. 2001;46(8):1908–1920.Google Scholar
Benz, R, Bauer, K. Permeation of hydrophilic molecules through the outer membrane of Gram-negative bacteria. European Journal of Biochemistry. 1988;176:119.Google Scholar
Burdige, DJ, Gardner, KG. Molecular weight distribution of dissolved organic carbon in marine sediment pore waters. Marine Chemistry. 1998;62(1–2):4564.CrossRefGoogle Scholar
Simpson, AJ, Kingery, WL, Hayes, MHB, Spraul, M, Humpfer, E, Dvortsak, P, et al. Molecular structures and associations of humic substances in the terrestrial environment. Naturwissenschaften. 2002;89(2):84–88.Google Scholar
Benner, R, Amon, RMW. The size–reactivity continuum of major bioelements in the ocean. Annual Review of Marine Science. 2015;7:185205.Google Scholar
Walker, BD, Beaupre, SR, Guilderson, TP, McCarthy, MD, Druffel, ERM. Pacific carbon cycling constrained by organic matter size, age and composition relationships. Nature Geoscience. 2016;9(12):888891.Google Scholar
Kelleher, BP, Simpson, AJ. Humic substances in soils: are they really chemically distinct? Environmental Science & Technology. 2006;40(15):4605–4611.CrossRefGoogle ScholarPubMed
Coolen, MJL, Overmann, J. Functional exoenzymes as indicators of metabolically active bacteria in 124,000-year-old sapropel layers of the eastern Mediterranean Sea. Applied and Environmental Microbiology. 2000;66(6):2589–2598.CrossRefGoogle ScholarPubMed
Coolen, MJL, Cypionka, H, Sass, AM, Sass, H, Overmann, J. Ongoing modification of Mediterranean Pleistocene sapropels mediated by prokaryotes. Science. 2002;296(5577):2407–2410.Google Scholar
Schmidt, JM. Microbial Extracellular Enzymes in Marine Sediments: Methods Development and Potential Activities in the Baltic Sea Deep Biosphere. Masters thesis, University of Tennessee, 2016.Google Scholar
Robador, A, Bruchert, V, Steen, AD, Arnosti, C. Temperature induced decoupling of enzymatic hydrolysis and carbon remineralization in long-term incubations of Arctic and temperate sediments. Geochimica et Cosmochimica Acta. 2010;74(8):2316–2326.CrossRefGoogle Scholar
Noinaj, N, Guillier, M, Barnard, TJ, Buchanan, SK. TonB-dependent transporters: regulation, structure, and function. Annual Review of Microbiology. 2010;64:4360.CrossRefGoogle ScholarPubMed
Arnosti, C, Repeta, DJ. Extracellular enzyme activity in anaerobic bacterial cultures – evidence of pullulanase activity among mesophilic marine bacteria. Applied and Environmental Microbiology. 1994;60(3):840–846.Google Scholar
Cuskin, F, Lowe, EC, Temple, MJ, Zhu, YP, Cameron, EA, Pudlo, NA, et al. Human gut Bacteroidetes can utilize yeast mannan through a selfish mechanism. Nature. 2015;517(7533):165–186.CrossRefGoogle ScholarPubMed
Tanoue, E, Handa, N. Distribution of particulate organic carbon and nitrogen in the Bering Sea and Northern North Pacific Ocean. Journal of the Oceanographical Society of Japan. 1979;35:4762.Google Scholar
Mayer, LM. Relationships between mineral surfaces and organic carbon concentrations in soils and sedimentsChemical Geology. 1994;114(3–4):347–363.Google Scholar
Mayer, LM. Surface area control of organic carbon accumulation in continental shelf sediments. Geochimica et Cosmochimica Acta. 1994;58(4):1271–1284.Google Scholar
Mayer, LM. Extent of coverage of mineral surfaces by organic matter in marine sediments. Geochimica et Cosmochimica Acta. 1999;63(2):207–215.Google Scholar
Knicker, H, Hatcher, PG. Survival of protein in an organic-rich sediment: possible protection by encapsulation in organic matter. Naturwissenschaften. 1997;84(6):231–234.Google Scholar
Moore, EK, Nunn, BL, Goodlett, DR, Harvey, HR. Identifying and tracking proteins through the marine water column: insights into the inputs and preservation mechanisms of protein in sediments. Geochimica et Cosmochimica Acta. 2012;83:324–359.Google Scholar
Wu, SC, Gschwend, PM. Sorption kinetics of hydrophobic organic compounds to natural sediments and soils. Environmental Science & Technology. 1986;20(7):717–725.Google Scholar
Espeland, E, Wetzel, R. Complexation, stabilization, and UV photolysis of extracellular and surface-bound glucosidase and alkaline phosphatase: implications for biofilm microbiota. Microbial Ecology. 2001;42(4):572–585.Google Scholar
Tietjen, T, Wetzel, R. Extracellular enzyme–clay mineral complexes: enzyme adsorption, alteration of enzyme activity, and protection from photodegradation. Aquatic Ecology. 2003;37(4):331–339.Google Scholar
Lalonde, K, Mucci, A, Ouellet, A, Gelinas, Y. Preservation of organic matter in sediments promoted by iron. Nature. 2012;483(7388):198200.Google Scholar
Barber, A, Brandes, J, Leri, A, Lalonde, K, Balind, K, Wirick, S, et al. Preservation of organic matter in marine sediments by inner-sphere interactions with reactive iron. Scientific Reports. 2017;7:366.Google Scholar
Cowie, GL, Hedges, JI. The role of anoxia in organic matter preservation in coastal sediments – relative stabilities of the major biochemicals under oxic and anoxic depositional conditions. Organic Geochemistry. 1992;19(1–3):229–234.Google Scholar
Canfield, DE. Factors influencing organic carbon preservation in marine sediments. Chemical Geology. 1994;114(3–4):315–329.Google Scholar
Keil, RG, Hu, FS, Tsamakis, EC, Hedges, JI. Pollen in marine sediments as an indicator of oxidation of organic matter. Nature. 1994;369(6482):639–641.Google Scholar
Hedges, JI, Hu, FS, Devol, AH, Hartnett, HE, Tsamakis, E, Keil, RG. Sedimentary organic matter preservation: a test for selective degradation under oxic conditions. American Journal of Science. 1999;299(7–9):529–555.Google Scholar
Kirk, TK, Farrell, RL. Enzymatic combustion – the microbial-degradation of lignin. Annual Review of Microbiology. 1987;41:465505.Google Scholar
Cowie, GL, Hedges, JI, Prahl, FG, Delange, GJ. Elemental and major biochemical changes across an oxidation front in an relict turbidite – an oxygen effect. Geochimica et Cosmochimica Acta. 1995;59(1):3346.Google Scholar
Cowie, GL, Calvert, SE, Pedersen, TF, Schulz, H, von Rad, U. Organic content and preservational controls in surficial shelf and slope sediments from the Arabian Sea (Pakistan margin). Marine Geology. 1999;161(1):2338.Google Scholar
Røy, H, et al. Aerobic microbial respiration in 86-million-year-old deep sea red clay. Science. 2012;336:922–925.CrossRefGoogle ScholarPubMed
Westrich, JT, Berner, RA. The role of sedimentary organic matter in bacterial sulfate reduction – the G model tested. Limnology and Oceanography. 1984;29(2):236–249.Google Scholar
Boudreau, BP, Ruddick, BR. On a reactive continuum representation of organic-matter diagenesis. American Journal of Science. 1991;291(5):507–538.Google Scholar
Tarutis, WJ. On the equivalence of the power and reactive continuum models of organic-matter diagenesis. Geochimica et Cosmochimica Acta. 1993;57(6):1349–1350.Google Scholar
Middelburg, JJ. A simple rate model for organic matter decomposition in marine sediments. Geochimica et Cosmochimica Acta. 1989;53(7):1577–1581.CrossRefGoogle Scholar
Hedges, JI, Eglinton, G, Hatcher, PG, Kirchman, DL, Arnosti, C, Derenne, S, et al. The molecularly-uncharacterized component of nonliving organic matter in natural environments. Organic Geochemistry. 2000;31(10):945–958.Google Scholar
Wieder, WR, Bonan, GB, Allison, SD. Global soil carbon projections are improved by modelling microbial processes. Nature Climate Change. 2013;3(10):909–912.Google Scholar
Bradley, JA, Amend, JP, LaRowe, DE. Bioenergetic controls on microbial ecophysiology in marine sediments. Frontiers in Microbiology. 2018;9:180.Google Scholar
Johnson, HP, Pruis, MJ. Fluxes of fluid and heat from the oceanic crustal reservoir. Earth and Planetary Science Letters. 2003;216(4):565–574.Google Scholar
Edmond, JM, Measures, C, McDuff, RE, Chan, LH, Collier, R, Grant, B, et al. Ridge crest hydrothermal activity and the balances of the major and minor elements in the ocean – Galapagos data. Earth and Planetary Science Letters. 1979;46(1):118.Google Scholar
Mottl, MJ, Wheat, CG. Hydrothermal circulation through midocean ridge flanks – fluxes of heat and magnesium. Geochimica et Cosmochimica Acta. 1994;58(10):2225–2237.Google Scholar
Elderfield, H, Schultz, A. Mid-ocean ridge hydrothermal fluxes and the chemical composition of the ocean. Annual Review of Earth and Planetary Sciences. 1996;24:191224.CrossRefGoogle Scholar
German, CR, Casciotti, KA, Dutay, JC, Heimburger, LE, Jenkins, WJ, Measures, CI, et al. Hydrothermal impacts on trace element and isotope ocean biogeochemistry. Philosophical Transactions of the Royal Society A –Mathematical Physical and Engineering Sciences. 2016;374(2081):20160035.Google Scholar
Stein, CA, Stein, S. Constraints on hydrothermal heat-flux through the oceanic lithosphere from global heat flow. Journal of Geophysical Research – Solid Earth. 1994;99(B2):3081–3095.Google Scholar
Alt, JC. Sulfur isotopic profile through the oceanic-crust – sulfur mobility and seawater–crustal sulfur exchange during hydrothermal alteration. Geology. 1995;23(7):585–588.Google Scholar
Summit, M, Baross, JA. A novel microbial habitat in the mid-ocean ridge subseafloor. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(5):2158–2163.Google Scholar
Santelli, CM, Orcutt, BN, Banning, E, Bach, W, Moyer, CL, Sogin, ML, et al. Abundance and diversity of microbial life in ocean crust. Nature. 2008;453(7195):653657.Google Scholar
Orcutt, BN, Sylvan, JB, Knab, NJ, Edwards, KJ. Microbial ecology of the dark ocean above, at, and below the seafloor. Microbiology and Molecular Biology Reviews. 2011;75(2):361422.Google Scholar
Sclater, JG, Jaupart, C, Galson, D. The heat flow through oceanic and continental crust and the heat loss of the Earth. Reviews of Geophysics. 1980;18(1):269311.Google Scholar
Butterfield, DA, McDuff, RE, Mottl, MJ, Lilley, MD, Lupton, JE, Massoth, GJ. Gradients in the composition of hydrothermal fluids from the Endeavor Segment vent field – phase-separation and brine loss. Journal of Geophysical Research – Solid Earth. 1994;99(B5):9561–9583.Google Scholar
Tivey, MK, Bradley, AM, Joyce, TM, Kadko, D. Insights into tide-related variability at seafloor hydrothermal vents from time-series temperature measurements. Earth and Planetary Science Letters. 2002;202(3–4):693707.Google Scholar
Lilley, MD, Butterfield, DA, Lupton, JE, Olson, EJ. Magmatic events can produce rapid changes in hydrothermal vent chemistry. Nature. 2003;422(6934):878–881.Google Scholar
Seewald, J, Cruse, A, Saccocia, P. Aqueous volatiles in hydrothermal fluids from the Main Endeavour Field, northern Juan de Fuca Ridge: temporal variability following earthquake activity. Earth and Planetary Science Letters. 2003;216(4):575–590.Google Scholar
Sarmiento, JL, Gruber, N. Ocean Biogeochemical Dynamics. Princeton, NJ: Princeton University Press, 2006.Google Scholar
Hansell, DA, Carlson, CA, Repeta, DJ, Schlitzer, R. Dissolved organic matter in the ocean a controversy stimulates new insights. Oceanography. 2009;22(4):202–211.CrossRefGoogle Scholar
Hansell, DA. Recalcitrant dissolved organic carbon fractions. Annual Review of Marine Science. 2013;5:421–445.Google Scholar
Druffel, ERM, Williams, PM, Bauer, JE, Ertel, Jr. Cycling of dissolved and particulate organic-matter in the open ocean. Journal of Geophysical Research – Oceans. 1992;97(C10):15639–15659.Google Scholar
Hertkorn, N, Benner, R, Frommberger, M, Schmitt-Kopplin, P, Witt, M, Kaiser, K, et al. Characterization of a major refractory component of marine dissolved organic matter. Geochimica et Cosmochimica Acta. 2006;70(12):29903010.Google Scholar
Aluwihare, LI, Repeta, DJ, Chen, RF. A major biopolymeric component to dissolved organic carbon in surface sea water. Nature. 1997;387(6629):166–169.CrossRefGoogle Scholar
Arakawa, N, Aluwihare, LI, Simpson, AJ, Soong, R, Stephens, BM, Lane-Coplen, D. Carotenoids are the likely precursor of a significant fraction of marine dissolved organic matter. Science Advances. 2017;3(9):e1602976.Google Scholar
Repeta, DJ. Chemical characterization and cycling of dissolved organic matter. In: Hansell, D, ed. Biogeochemistry of Marine Dissolved Organic Matter. Washington, DC: American Association for the Advancement of Science, 2015, pp. 2163.Google Scholar
Campbell, AC, Palmer, MR, Klinkhammer, GP, Bowers, TS, Edmond, JM, Lawrence, JR, et al. Chemistry of hot springs on the Mid-Atlantic Ridge. Nature. 1988;335(6190):514–519.CrossRefGoogle Scholar
Conway, TM, John, SG. Quantification of dissolved iron sources to the North Atlantic Ocean. Nature. 2014;511(7508):212–215.Google Scholar
Resing, JA, Sedwick, PN, German, CR, Jenkins, WJ, Moffett, JW, Sohst, BM, et al. Basin-scale transport of hydrothermal dissolved metals across the South Pacific Ocean. Nature. 2015;523(7559):200–203.Google Scholar
McCollom, TM. Observational, experimental, and theoretical constraints on carbon cycling in mid-ocean ridge hydrothermal systems. In: Lowell, RP, Seewald, JS, Metaxas, A, Perfit, MR, eds. Magma to Microbe: Modeling Hydrothermal Processes at Ocean Spreading Centers. Washington, DC: American Geophysical Union, 2008. pp. 193213.Google Scholar
Pineau, F, Javoy, M. Carbon isotopes and concentrations in mid-ocean ridge basalts. Earth and Planetary Science Letters. 1983;62(2):239–257.CrossRefGoogle Scholar
DesMarais, DJ, Moore, JG. Carbon and its isotopes in mid-ocean basaltic glasses. Earth and Planetary Science Letters. 1984;69(1):4357.Google Scholar
Kroopnick, PM. The distribution of C-13 of sigma-CO2 in the world ocean. Deep-Sea Research Part A – Oceanographic Research Papers. 1985;32(1):5784.Google Scholar
Proskurowski, G, Lilley, MD, Brown, TA. Isotopic evidence of magmatism and seawater bicarbonate removal at the endeavour hydrothermal system. Earth and Planetary Science Letters. 2004;225(1–2):5361.Google Scholar
Alt, JC, Teagle, DAH. Hydrothermal alteration of upper oceanic crust formed at a fast-spreading ridge: mineral, chemical, and isotopic evidence from ODP Site 801. Chemical Geology. 2003;201(3–4):191211.Google Scholar
Eickmann, B, Bach, W, Rosner, M, Peckmann, J. Geochemical constraints on the modes of carbonate precipitation in peridotites from the Logatchev Hydrothermal Vent Field and Gakkel Ridge. Chemical Geology. 2009;268(1–2):97106.Google Scholar
Beaupre, SR, Druffel, ERM. Constraining the propagation of bomb-radiocarbon through the dissolved organic carbon (DOC) pool in the northeast Pacific Ocean. Deep-Sea Research Part I – Oceanographic Research Papers. 2009;56(10):1717–1726.Google Scholar
Reeves, EP, McDermott, JM, Seewald, JS. The origin of methanethiol in midocean ridge hydrothermal fluids. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(15):5474–5479.Google Scholar
Riedel, M, Collett, TS, Malone, M, Scientists, E. Expedition 311 synthesis: scientific findings. Proceedings of the Integrated Ocean Drilling Program. 2006;311:2.Google Scholar
Von Damm, KL, Lilley, MD. Diffuse flow hydrothermal fluids from 9-degrees 50’ N East Pacific Rise: origin, evolution and biogeochemical controls. In: William SDW, Edward FD, Deborah SK, John AB, Craig SC, eds. The Subseafloor Biosphere at Mid-Ocean Ridges. Geophysical Monograph Series 144. Washington, DC: American Geophysical Union, 2004, pp. 245–268.Google Scholar
Proskurowski, G, Lilley, MD, Seewald, JS, Früh-Green, GL, Olson, EJ, Lupton, JE, et al. Abiogenic hydrocarbon production at Lost City hydrothermal field. Science. 2008;319(5863):604–607.Google Scholar
Yucel, M, Luther, GW. Temporal trends in vent fluid iron and sulfide chemistry following the 2005/2006 eruption at East Pacific Rise, 9 degrees 50’ N. Geochemistry, Geophysics, Geosystems. 2013;14(4):759–765.Google Scholar
Butterfield, DA, Massoth, GJ, McDuff, RE, Lupton, JE, Lilley, MD. Geochemistry of hydrothermal fluids from Axial Seamount hydrothermal emissions study vent field, Juan de Fuca Ridge – subseafloor boiling and subsequent fluid-rock interactionJournal of Geophysical Research – Solid Earth and Planets. 1990;95(B8):12895–12921.Google Scholar
Butterfield, DA, Roe, KK, Lilley, MD, Huber, JA, Baross, JA, Embley, RW, et al. Mixing, reaction and microbial activity in the sub-seafloor revealed by temporal and spatial variation in diffuse flow vents at Axial Volcano. Subseafloor Biosphere at Mid-Ocean Ranges. 2004;144:269–289.Google Scholar
Lang, SQ, Butterfield, DA, Lilley, MD, Johnson, HP, Hedges, JI. Dissolved organic carbon in ridge-axis and ridge-flank hydrothermal systems. Geochimica et Cosmochimica Acta. 2006;70(15):3830–3842.CrossRefGoogle Scholar
Charlou, JL, Donval, JP. Hydrothermal methane venting between 12-degrees-N and 26-degrees-N along the Mid-Atlantic Ridge. Journal of Geophysical Research – Solid Earth. 1993;98(B6):9625–9642.Google Scholar
Charlou, JL, Donval, JP, Jean-Baptiste, P, Dapoigny, A, Rona, PA. Gases and helium isotopes in high temperature solutions sampled before and after ODP Leg 158 drilling at TAG hydrothermal field (26N MAR). Geophysical Research Letters. 1996;23:3491–3494.Google Scholar
Charlou, JL, Donval, JP, Konn, C, Ondreas, H, Fouquet, Y, Jean-Baptiste, P, et al. High production and fluxes of H2 and CH4 and evidence of abiotic hydrocarbon synthesis by serpentinization in ultramafic-hosted hydrothermal systems on the Mid-Atlantic Ridge. In: Rona, PA, ed. Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. Geophysical Monograph Series 188. Washingon, DC: American Geophysical Union, 2010, pp. 265–296.Google Scholar
Proskurowski, G, Lilley, MD, Olson, EJ. Stable isotopic evidence in support of active microbial methane cycling in low-temperature diffuse flow vents at 9 degrees 50’ N East Pacific Rise. Geochimica et Cosmochimica Acta. 2008;72(8):2005–2023.Google Scholar
McCarthy, M, Beaupre, S, Walker, B, Voparil, I, Guilderson, T, Druffel, E. Chemosynthetic origin of C-14-depleted dissolved organic matter in a ridge-flank hydrothermal system. Nature Geoscience. 2011;4(1):32–36.Google Scholar
Haberstroh, PR, Karl, DM. Dissolved free amino-acids in hydrothermal vent habitats of the Guaymas Basin. Geochimica et Cosmochimica Acta. 1989;53(11):2937–2945.Google Scholar
Von Damm, KL. Seafloor hydrothermal activity: black smoker chemistry and chimneys. Annual Review of Earth and Planetary Sciences. 1990;18:173204.Google Scholar
Lin, HT, Amend, JP, LaRowe, DE, Bingham, JP, Cowen, JP. Dissolved amino acids in oceanic basaltic basement fluids. Geochimica et Cosmochimica Acta. 2015;164:175–190.Google Scholar
Lin, Y, Koch, B, Feseker, T, Ziervogel, K, Goldhammer, T, Schmidt, F, et al. Near-surface heating of young rift sediment causes mass production and discharge of reactive dissolved organic matter. Scientific Reports. 2017;7:44864.Google Scholar
McKay, L, Klokman, VW, Mendlovitz, HP, LaRowe, DE, Hoer, DR, Albert, D, et al. Thermal and geochemical influences on microbial biogeography in the hydrothermal sediments of Guaymas Basin, Gulf of California. Environmental Microbiology Reports. 2016;8(1):150–161.Google Scholar
Cruse, AM, Seewald, JS. Geochemistry of low-molecular weight hydrocarbons in hydrothermal fluids from Middle Valley, northern Juan de Fuca Ridge. Geochimica et Cosmochimica Acta. 2006;70(8):2073–2092.Google Scholar
Wheat, C, Mottl, M, Fisher, A, Kadko, D, Davis, E, Baker, E. Heat flow through a basaltic outcrop on a sedimented young ridge flank. Geochemistry, Geophysics, Geosystems. 2004;5:Q12006.CrossRefGoogle Scholar
Wheat, CG, Hulme, SM, Fisher, AT, Orcutt, BN, Becker, K. Seawater recharge into oceanic crust: IODP Exp 327 Site U1363 Grizzly Bare outcrop. Geochemistry, Geophysics, Geosystems. 2013;14(6):1957–1972.Google Scholar
Walker, BD, McCarthy, MD, Fisher, AT, Guilderson, TP. Dissolved inorganic carbon isotopic composition of low-temperature axial and ridge-flank hydrothermal fluids of the Juan de Fuca Ridge. Marine Chemistry. 2008;108(1–2):123–136.Google Scholar
Lin, HT, Cowen, JP, Olson, EJ, Amend, JP, Lilley, MD. Inorganic chemistry, gas compositions and dissolved organic carbon in fluids from sedimented young basaltic crust on the Juan de Fuca Ridge flanks. Geochimica et Cosmochimica Acta. 2012;85:213–227.Google Scholar
Lin, HT, Cowen, JP, Olson, EJ, Lilley, MD, Jungbluth, SP, Wilson, ST, et al. Dissolved hydrogen and methane in the oceanic basaltic biosphere. Earth and Planetary Science Letters. 2014;405:6273.Google Scholar
Lin, H-T, Repeta, DJ, Xu, L, Rappe, MS. Selective removal of isotopically enriched dissolved organic carbon from basalt-hosted deep subseafloor fluids of the Juan de Fuca Ridge. Earth and Planetary Science Letters. 2019;513:156–165.Google Scholar
Meyer, JL, Jaekel, U, Tully, BJ, Glazer, BT, Wheat, CG, Lin, HT, et al. A distinct and active bacterial community in cold oxygenated fluids circulating beneath the western flank of the Mid-Atlantic ridge. Scientific Reports. 2016;6:22541.CrossRefGoogle ScholarPubMed
Walter, SRS, Jaekel, U, Osterholz, H, Fisher, AT, Huber, JA, Pearson, A, et al. Microbial decomposition of marine dissolved organic matter in cool oceanic crust. Nature Geoscience. 2018;11(5):334–339.Google Scholar
McDermott, JM, Seewald, JS, German, CR, Sylva, SP. Pathways for abiotic organic synthesis at submarine hydrothermal fields. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(25):7668–7672.Google Scholar
Douville, E, Charlou, JL, Oelkers, EH, Bienvenu, P, Colon, CFJ, Donval, JP, et al. The rainbow vent fluids (36 degrees 14’ N, MAR): the influence of ultramafic rocks and phase separation on trace metal content in Mid-Atlantic Ridge hydrothermal fluids. Chemical Geology. 2002;184(1–2):3748.Google Scholar
Charlou, JL, Donval, JP, Fouquet, Y, Jean-Baptiste, P, Holm, N. Geochemistry of high H2 and CH4 vent fluids issuing from ultramafic rocks at the Rainbow hydrothermal field (36 degrees 14’ N, MAR). Chemical Geology. 2002;191(4):345–359.Google Scholar
Jean-Baptiste, P, Fourre, E, Charlou, JL, German, CR, Radford-Knoery, J. Helium isotopes at the Rainbow hydrothermal site (Mid-Atlantic Ridge, 36 degrees 14’ N). Earth and Planetary Science Letters. 2004;221(1–4):325–335.Google Scholar
Kelley, DS, Karson, JA, Früh-Green, GL, Yoerger, DR, Shank, TM, Butterfield, DA, et al. A serpentinite-hosted ecosystem: the Lost City hydrothermal field. Science. 2005;307(5714):1428–1434.Google Scholar
Lang, SQ, Butterfield, DA, Schulte, M, Kelley, DS, Lilley, MD. Elevated concentrations of formate, acetate and dissolved organic carbon found at the Lost City hydrothermal field. Geochimica et Cosmochimica Acta. 2010;74(3):941–952.Google Scholar
Lang, SQ, Früh-Green, GL, Bernasconi, SM, Butterfield, DA. Sources of organic nitrogen at the serpentinite-hosted Lost City hydrothermal field. Geobiology. 2013;11(2):154–169.Google Scholar
Seyfried, WE, Pester, NJ, Tutolo, BM, Ding, K. The Lost City hydrothermal system: constraints imposed by vent fluid chemistry and reaction path models on subseafloor heat and mass transfer processes. Geochimica et Cosmochimica Acta. 2015;163:5979.Google Scholar
Reeves, EP, Seewald, JS, Saccocia, P, Bach, W, Craddock, PR, Shanks, WC, et al. Geochemistry of hydrothermal fluids from the PACMANUS, Northeast Pual and Vienna Woods hydrothermal fields, Manus Basin, Papua New Guinea. Geochimica et Cosmochimica Acta. 2011;75(4):1088–1123.Google Scholar
Sakai, H, Gamo, T, Kim, ES, Tsutsumi, M, Tanaka, T, Ishibashi, J, et al. Venting of carbon-dioxide rich fluid and hydrate formation in Mid-Okinawa Trough Backarc Basin. Science. 1990;248(4959):1093–1096.Google Scholar
Ishibashi, J, Sano, Y, Wakita, H, Gamo, T, Tsutsumi, M, Sakai, H. Helium and carbon geochemistry of hydrothermal fluids from the Mid-Okinawa Trough Back-arc Basin, southwest of Japan. Chemical Geology. 1995;123(1–4):115.Google Scholar
Welhan, JA, Craig, H. Methane, hydrogen and helium in hydrothermal fluids at 21N on the East Pacific Rise. In: Rona, PA, ed. Hydrothermal Processes at Seafloor Spreading Centers. New York: Plenum, 1983, pp. 391409.Google Scholar
Lilley, MD, Baross, JA, Gordon, LI. Reduced gases and bacteria in hydrothermal fluids: the Galapagos Spreading Center and 21 deg N East Pacific Rise. In: Rona, PA, Bostrom, K, Leubier, L, Smith, KL Jr., eds. Hydrothermal Processes at Seafloor Spreading Centers. New York: Plenum, 1983, pp. 411–449.Google Scholar
Evans, WC, White, LD, Rapp, JB. Geochemistry of some gases in hydrothermal fluids from the southern Juan-de-Fuca-Ridge. Journal of Geophysical Research – Solid Earth and Planets. 1988;93(B12):15305–15313.Google Scholar
Merlivat, L, Pineau, F, Javoy, M. Hydrothermal vent waters at 13-degrees-N on the East Pacific Rise – isotopic composition and gas concentration. Earth and Planetary Science Letters. 1987;84(1):100–108.CrossRefGoogle Scholar
Jeanbaptiste, P, Charlou, JL, Stievenard, M, Donval, JP, Bougault, H, Mevel, C. Helium and methane measurements in hydrothermal fluids from the Mid-Atlantic Ridge – the Snake Pit Site at 23-degrees-N. Earth and Planetary Science Letters. 1991;106(1–4):1728.Google Scholar
James, RH, Elderfield, H, Palmer, MR. The chemistry of hydrothermal fluids from the Broken Spur Site, 29-degrees-N Mid-Atlantic Ridge. Geochimica et Cosmochimica Acta. 1995;59(4):651–659.Google Scholar
Lein, AY, Grichuk, DV, Gurvich, EG, Bogdanov, YA. A new type of hydrogen- and methane-rich hydrothermal solutions in the rift zone of the Mid-Atlantic Ridge. Doklady Earth Sciences. 2000;375(9):1391–1394.Google Scholar
Martens, CS. Generation of short chain organic-acid anions in hydrothermally altered sediments of the Guaymas Basin, Gulf of California. Applied Geochemistry. 1990;5:71–76.Google Scholar
Hawkes, JA, Rossel, PE, Stubbins, A, Butterfield, D, Connelly, DP, Achterberg, EP, et al. Efficient removal of recalcitrant deep-ocean dissolved organic matter during hydrothermal circulation. Nature Geoscience. 2015;8(11):856–860.Google Scholar
Wellsbury, P, Goodman, K, Barth, T, Cragg, B, Barnes, S, Parkes, R. Deep marine biosphere fuelled by increasing organic matter availability during burial and heating. Nature. 1997;388(6642):573–576.Google Scholar
Seewald, JS, Seyfried, WE, Thornton, EC. Organic-rich sediment alteration – an experimental and theoretical study at elevated temperatures and pressures. Applied Geochemistry. 1990;5:193209.Google Scholar
McCollom, TM, Shock, EL. Geochemical constraints on chemolithoautotrophic metabolism by microorganisms in seafloor hydrothermal systems. Geochimica et Cosmochimica Acta. 1997;61(20):4375–4391.Google Scholar
Takai, K, Nakamura, K, Toki, T, Tsunogai, U, Miyazaki, M, Miyazaki, J, et al. Cell proliferation at 122 degrees C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(31):10949–10954.Google Scholar
Karl, DM, Wirsen, CO, Jannasch, HW. Deep-sea primary production at the Galapagos hydrothermal vents. Science. 1980;207(4437):1345–1347.Google Scholar
Edwards, KJ, Becker, K, Colwell, F. The deep, dark energy biosphere: intraterrestrial life on Earth. Annual Review of Earth and Planetary Sciences. 2012;40:551–568.CrossRefGoogle Scholar
Wankel, SD, Germanovich, LN, Lilley, MD, Genc, G, DiPerna, CJ, Bradley, AS, et al. Influence of subsurface biosphere on geochemical fluxes from diffuse hydrothermal fluids. Nature Geoscience. 2011;4(7):461–468.Google Scholar
Mottl, MJ. Partitioning of energy and mass fluxes between mid-ocean ridge axes and flanks at high and low temperature. In: Halbach, PE, Tunnicliffe, V, Hein, JR, eds. Energy and Mass Transfer in Marine Hydrothermal Systems. Berlin: Dahlem University Press, 2003, pp. 271–286.Google Scholar
Wheat, CG, Fisher, AT, McManus, J, Hulme, SM, Orcutt, BN. Cool seafloor hydrothermal springs reveal global geochemical fluxes. Earth and Planetary Science Letters. 2017;476:179–188.Google Scholar
Elderfield, H, Wheat, CG, Mottl, MJ, Monnin, C, Spiro, B. Fluid and geochemical transport through oceanic crust: a transect across the eastern flank of the Juan de Fuca Ridge. Earth and Planetary Science Letters. 1999;172(1–2):151–165.Google Scholar
Mottl, MJ, Wheat, G, Baker, E, Becker, N, Davis, E, Feely, R, et al. Warm springs discovered on 3.5 Ma oceanic crust, eastern flank of the Juan de Fuca Ridge. Geology. 1998;26(1):51–54.Google Scholar
Sansone, FJ, Mottl, MJ, Olson, EJ, Wheat, CG, Lilley, MD. CO2-depleted fluids from mid-ocean ridge-flank hydrothermal springs. Geochimica et Cosmochimica Acta. 1998;62(13):2247–2252.Google Scholar
Maloszewski, P, Zuber, A. Influence of matrix diffusion and exchange-reactions on radiocarbon ages in fissured carbonate aquifers. Water Resources Research. 1991;27(8):1937–1945.Google Scholar
Bethke, CM, Johnson, TM. Paradox of groundwater age. Geology. 2002;30(2):107–110.Google Scholar
Bethke, CM, Johnson, TM. Groundwater age and groundwater age dating. Annual Review of Earth and Planetary Sciences. 2008;36:121–152.Google Scholar
Rona, PA, Widenfalk, L, Bostrom, K. Serpentinized ultramafics and hydrothermal activity at the Mid-Atlantic Ridge crest near 15-degrees-N. Journal of Geophysical Research – Solid Earth and Planets. 1987;92(B2):1417–1427.Google Scholar
Charlou, JL, Bougault, H, Appriou, P, Jeanbaptiste, P, Etoubleau, J, Birolleau, A. Water column anomalies associated with hydrothermal activity between 11-degrees-40’ and 13-degrees-N on the east pacific rise – discrepancies between tracers. Deep-Sea Research Part A – Oceanographic Research Papers. 1991;38(5):569–596.Google Scholar
Rona, PA, Bougault, H, Charlou, JL, Appriou, P, Nelsen, TA, Trefry, JH, et al. Hydrothermal circulation, serpentinization, and degassing at a rift-valley fracture-zone intersection – Mid-Atlantic Ridge near 15-degrees-N, 45-degrees-W. Geology. 1992;20(9):783–786.Google Scholar
Schrenk, MO, Brazelton, WJ, Lang, SQ. Serpentinization, carbon, and deep life. In: Hazen, RM, Jones, AP, Baross, JA, eds. Carbon in Earth. Reviews in Mineralogy & Geochemistry. 75. Chantilly: Mineralogical Society of America, 2013, pp. 575606.Google Scholar
German, CR, Bowen, A, Coleman, ML, Honig, DL, Huber, JA, Jakuba, MV, et al. Diverse styles of submarine venting on the ultraslow spreading Mid-Cayman Rise. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(32):14020–14025.Google Scholar
Haggerty, JA. Evidence from fluid seeps atop serpentine seamounts in the mariana fore-arc – clues for emplacement of the seamounts and their relationship to fore-arc tectonics. Marine Geology. 1991;102(1–4):293309.Google Scholar
Kelley, DS, Karson, JA, Blackman, DK, Früh-Green, GL, Butterfield, DA, Lilley, MD, et al. An off-axis hydrothermal vent field near the Mid-Atlantic Ridge at 30 degrees N. Nature. 2001;412(6843):145–149.Google Scholar
Monnin, C, Chavagnac, V, Boulart, C, Menez, B, Gerard, M, Gerard, E, et al. Fluid chemistry of the low temperature hyperalkaline hydrothermal system of Prony Bay (New Caledonia). Biogeosciences. 2014;11(20):5687–5706.Google Scholar
Ohara, Y, Reagan, MK, Fujikura, K, Watanabe, H, Michibayashi, K, Ishii, T, et al. A serpentinite-hosted ecosystem in the Southern Mariana Forearc. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(8):2831–2835.Google Scholar
Konn, C, Charlou, JL, Holm, NG, Mousis, O. The production of methane, hydrogen, and organic compounds in ultramafic-hosted hydrothermal vents of the Mid-Atlantic Ridge. Astrobiology. 2015;15(5):381–399.Google Scholar
Bonatti, E, Lawrence, JR, Hamlyn, PR, Breger, D. Aragonite from deep-sea ultramafic rocks. Geochimica et Cosmochimica Acta. 1980;44(8):1207–1214.Google Scholar
Keir, RS. A note on the fluxes of abiogenic methane and hydrogen from mid-ocean ridges. Geophysical Research Letters. 2010;37:L24609.Google Scholar
McCollom, TM, Seewald, JS. Abiotic synthesis of organic compounds in deep-sea hydrothermal environments. Chemical Reviews. 2007;107(2):382401.Google Scholar
Wang, DT, Reeves, EP, McDermott, JM, Seewald, JS, Ono, S. Clumped isotopologue constraints on the origin of methane at seafloor hot springs. Geochimica et Cosmochimica Acta. 2018;223:141–158.Google Scholar
Kelley, DS, Früh-Green, GL. Abiogenic methane in deep-seated mid-ocean ridge environments: insights from stable isotope analyses. Journal of Geophysical Research – Solid Earth. 1999;104(B5):10439–10460.Google Scholar
Kelley, DS, Früh-Green, GL. Volatile lines of descent in submarine plutonic environments: insights from stable isotope and fluid inclusion analyses. Geochimica et Cosmochimica Acta. 2001;65(19):3325–3346.Google Scholar
Bradley, AS, Summons, RE. Multiple origins of methane at the Lost City hydrothermal field. Earth and Planetary Science Letters. 2010;297(1–2):3441.Google Scholar
Pisapia, C, Gerard, E, Gerard, M, Lecourt, L, Lang, SQ, Pelletier, B, et al. Mineralizing filamentous bacteria from the Prony Bay hydrothermal field give new insights into the functioning of serpentinization-based subseafloor ecosystems. Frontiers in Microbiology. 2017;8:18.Google Scholar
Lang, SQ, Früh-Green, GL, Bernasconi, SM, Brazelton, WJ, Schrenk, MO, McGonigle, JM. Deeply-sourced formate fuels sulfate reducers but not methanogens at Lost City hydrothermal field. Scientific Reports. 2018;8:755.Google Scholar
Kawka, OE, Simoneit, BRT. Survey of hydrothermally-generated petroleums from the Guaymas Basin spreading center. Organic Geochemistry. 1987;11(4):311–328.Google Scholar
Kadko, D, Baross, J, Alt, J. The magnitude and global implications of hydrothermal flux. In: Humphris, SE, Zierenberg, RA, Mullineaux, LS, Thomson, RE, eds. Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions. Geophysical Monograph Series. Washington, DC: American Geophysical Union, 1995, pp. 446–466.Google Scholar
Saal, AE, Hauri, EH, Langmuir, CH, Perfit, MR. Vapour undersaturation in primitive mid-ocean-ridge basalt and the volatile content of Earth’s upper mantle. Nature. 2002;419(6906):451–455.Google Scholar
Staudigel, H, Hart, SR, Schmincke, HU, Smith, BM. Cretaceous ocean crust at DSDP Site-417 and Site-418 – carbon uptake from weathering versus loss by magmatic outgassing. Geochimica et Cosmochimica Acta. 1989;53(11):3091–3094.Google Scholar
Alt, JC, Teagle, DAH. The uptake of carbon during alteration of ocean crust. Geochimica et Cosmochimica Acta. 1999;63(10):1527–1535.Google Scholar
Schwarzenbach, EM, Lang, SQ, Frah-Green, GL, Lilley, MD, Bemasconi, SM, Mehay, S. Sources and cycling of carbon in continental, serpentinite-hosted alkaline springs in the Voltri Massif, Italy. Lithos. 2013;177:226–244.Google Scholar
Alt, JC, Schwarzenbach, EM, Frueh-Green, GL, Shanks, WC, Bernasconi, SM, Garrido, CJ, et al. The role of serpentinites in cycling of carbon and sulfur: seafloor serpentinization and subduction metamorphism. Lithos. 2013;178:4054.Google Scholar
Simoneit, BRT, Lonsdale, PF. Hydrothermal petroleum in mineralized mounds at the seabed of Guaymas Basin. Nature. 1982;295(5846):198202.Google Scholar
Kvenvolden, KA, Simoneit, BRT. Hydrothermally derived petroleum – examples from Guaymas Basin, Gulf of California, and Escanaba Trough, northeast Pacific-Ocean. AAPG Bulletin – American Association of Petroleum Geologists. 1990;74(3):223–237.Google Scholar
Tissot, BP, Welte, DH. Petroleum Formation and Occurrence. Berlin: Springer, 1984.Google Scholar
Tissot, B, Espitalie, J. Thermal evolution of organic-matter in sediments – application of a mathematical simulation – petroleum potential of sedimentary basins and reconstructing thermal history of sediments. Revue De L Institut Francais Du Petrole. 1975;30(5):743–777.Google Scholar
Lewan, MD, Winters, JC, McDonald, JH. Generation of oil-like pyrolyzates from organic-rich shales. Science. 1979;203(4383):897–899.Google Scholar
Seewald, JS. Aqueous geochemistry of low molecular weight hydrocarbons at elevated temperatures and pressures: constraints from mineral buffered laboratory experiments. Geochimica et Cosmochimica Acta. 2001;65(10):1641–1664.Google Scholar
Shock, EL, Canovas, P, Yang, ZM, Boyer, G, Johnson, K, Robinson, K, et al. Thermodynamics of organic transformations in hydrothermal fluids. Thermodynamics of Geothermal Fluids. 2013;76:311–350.Google Scholar
Simoneit, BRT. Hydrothermal petroleum – genesis, migration, and deposition in Guaymas Basin, Gulf of California. Canadian Journal of Earth Sciences. 1985;22(12):1919–1929.Google Scholar
Didyk, BM, Simoneit, BRT. Hydrothermal oil of Guaymas Basin and implications for petroleum formation mechanisms. Nature. 1989;342(6245):65–69.Google Scholar
Rushdi, AI, Simoneit, BRT. Condensation reactions and formation of amides, esters, and nitriles under hydrothermal conditions. Astrobiology. 2004;4(2):211–224.Google Scholar
Simoneit, B, Lein, A, Peresypkin, V, Osipov, G. Composition and origin of hydrothermal petroleum and associated lipids in the sulfide deposits of the Rainbow Field (Mid-Atlantic Ridge at 36 degrees N). Geochimica et Cosmochimica Acta. 2004;68(10):2275–2294.Google Scholar
Parkes, RJ, Taylor, J, Jorckramberg, D. Demonstration, using Desulfobacter sp., of 2 pools of acetate with different biological availabilities in marine pore water. Marine Biology. 1984;83(3):271–276.Google Scholar
Simoneit, BRT, Kawka, OE, Brault, M. Origin of gases and condensates in the Guaymas Basin hydrothermal system (Gulf of California). Chemical Geology. 1988;71(1–3):169–182.Google Scholar
Murdoch, LC, Germanovich, LN, Wang, H, Onstott, TC, Elsworth, D, Stetler, L, et al. Hydrogeology of the vicinity of Homestake mine, South Dakota, USA. Hydrogeology Journal. 2012;20(1):2743.Google Scholar
Holland, G, Lollar, BS, Li, L, Lacrampe-Couloume, G, Slater, GF, Ballentine, CJ. Deep fracture fluids isolated in the crust since the Precambrian era. Nature. 2013;497(7449):357–360.Google Scholar
Pedersen, K. Exploration of deep intraterrestrial microbial life: current perspectives. FEMS Microbiology Letters. 2000;185(1):916.Google Scholar
Colwell, FS, D’Hondt, S. Nature and extent of the deep biosphere. Carbon in Earth. 2013;75:547–574.Google Scholar
Huber, H, Huber, R, Lüdemann, H-D, Stetter, KO. Search for hyperhermophilic microorganisms in fluids obtained from the KTB pump test. Scientific Drilling. 1994;4:127–129.Google Scholar
Whitman, WB, Coleman, DC, Wiebe, WJ. Prokaryotes: the unseen majority. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(12):6578–6583.Google Scholar
Kallmeyer, J, Pockalny, R, Adhikari, RR, Smith, DC, D’Hondt, S. Global distribution of microbial abundance and biomass in subseafloor sediment. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(40):16213–16216.Google Scholar
McMahon, S, Parnell, J. Weighing the deep continental biosphere. FEMS Microbiology Ecology. 2014;87(1):113–120.Google Scholar
Fredrickson, JK, Onstott, TC, eds. Biogeochemical and Geological Significance of Subsurface Microbiology. New York: Wiley-Liss, 2001.Google Scholar
Sahl, JW, Schmidt, RH, Swanner, ED, Mandernack, KW, Templeton, AS, Kieft, TL, et al. Subsurface microbial diversity in deep-granitic-fracture water in Colorado. Applied and Environmental Microbiology. 2008;74(1):143–152.Google Scholar
Chivian, D, Brodie, EL, Alm, EJ, Culley, DE, Dehal, PS, DeSantis, TZ, et al. Environmental genomics reveals a single-species ecosystem deep within earth. Science. 2008;322(5899):275–278.Google Scholar
Osburn, MR, LaRowe, DE, Momper, LM, Amend, JP. Chemolithotrophy in the continental deep subsurface: Sanford Underground Research Facility (SURF), USA. Frontiers in Microbiology. 2014;5:610.Google Scholar
Probst, AJ, Castelle, CJ, Singh, A, Brown, CT, Anantharaman, K, Sharon, I, Hug, LA, Burstein, D, Emerson, JB, Thomas, BC, Banfield, JF. Genomic resolution of a cold subsurface aquifer community provides metabolic insights for novel microbes adapted to high CO2 concentrations. Environmental Microbiology. 2016;19:459–474.Google Scholar
Stevens, TO, McKinley, JP. Lithoautotrophic microbial ecosystems in deep basalt aquifers. Science. 1995;270(5235):450–454.Google Scholar
Veto, I, Futo, I, Horvath, I, Szanto, Z. Late and deep fermentative methanogenesis as reflected in the H–C–O–S isotopy of the methane-water system in deep aquifers of the Pannonian Basin (SE Hungary). Organic Geochemistry. 2004;35(6):713–723.Google Scholar
Fredrickson, JK, Balkwill, DL. Geomicrobial processes and biodiversity in the deep terrestrial subsurface. Geomicrobiology Journal. 2006;23(6):345–356.Google Scholar
Lapworth, DJ, Gooddy, DC, Butcher, AS, Morris, BL. Tracing groundwater flow and sources of organic carbon in sandstone aquifers using fluorescence properties of dissolved organic matter (DOM). Applied Geochemistry. 2008;23(12):3384–3390.Google Scholar
Vetter, A, Mangelsdorf, K, Wolfgramm, M, Rauppach, K, Schettler, G, Vieth-Hillebrand, A. Variations in fluid chemistry and membrane phospholipid fatty acid composition of the bacterial community in a cold storage groundwater system during clogging events. Applied Geochemistry. 2012;27(6):1278–1290.Google Scholar
Grundger, F, Jimenez, N, Thielemann, T, Straaten, N, Luders, W, Richnow, HH, et al. Microbial methane formation in deep aquifers of a coal-bearing sedimentary basin, Germany. Frontiers in Microbiology. 2015;6:200.Google Scholar
Schloemer, S, Elbracht, J, Blumenberg, M, Illing, CJ. Distribution and origin of dissolved methane, ethane and propane in shallow groundwater of Lower Saxony, Germany. Applied Geochemistry. 2016;67:118–132.Google Scholar
Bell, RA, Darling, WG, Ward, RS, Basava-Reddi, L, Halwa, L, Manamsa, K, et al. A baseline survey of dissolved methane in aquifers of Great Britain. Science of the Total Environment. 2017;601:1803–1813.Google Scholar
Murphy, EM, Schramke, JA, Fredrickson, JK, Bledsoe, HW, Francis, AJ, Sklarew, DS, et al. The influence of microbial activity and sedimentary organic-carbon on the isotope geochemistry of the middendorf aquifer. Water Resources Research. 1992;28(3):723–740.Google Scholar
Means, JL, Hubbard, N. Short-chain aliphatic acid anions in deep subsurface brines – a review of their origin, occurrence, properties, and importance of new data on their distribution and geochemical implications in the paleo-duro-basin, Texas. Organic Geochemistry. 1987;11(3):177–191.Google Scholar
Lundegard, PD, Kharaka, YK, eds. Distribution and Occurrence of Organic Acids in Subsurface Waters. Berlin: Springer, 1994.Google Scholar
Aitken, CM, Jones, DM, Larter, SR. Anaerobic hydrocarbon biodegradation in deep subsurface oil reservoirs. Nature. 2004;431(7006):291–294.Google Scholar
Strapoc, D, Mastalerz, M, Dawson, K, Macalady, J, Callaghan, AV, Wawrik, B, et al. Biogeochemistry of microbial coal-bed methane. Annual Review of Earth and Planetary Sciences. 2011;39:617–656.Google Scholar
Pedersen, K. Diversity and activity of microorganisms in deep igneous rock aquifers of the Fennoscandian Shield. In: Fredrickson, JK, Fletcher, M, eds. Subsurface Microbiology and Biogeochemistry, New York: Wiley-Liss, 2001, pp. 97139.Google Scholar
Amend, JP, Teske, A. Expanding frontiers in deep subsurface microbiology. Palaeogeography, Palaeoclimatology, Palaeoecology. 2005;219(1–2):131–155.Google Scholar
Lollar, BS, Onstott, TC, Lacrampe-Couloume, G, Ballentine, CJ. The contribution of the Precambrian continental lithosphere to global H2 production. Nature. 2014;516(7531):379–382.Google Scholar
Boston, PJ, Spilde, MN, Northup, DE, Melim, LA, Soroka, DS, Kleina, LG, et al. Cave biosignature suites: microbes, minerals, and Mars. Astrobiology. 2001;1(1):2555.Google Scholar
Northup, DE, Lavoie, KH. Geomicrobiology of caves: a review. Geomicrobiology Journal. 2001;18(3):199222.Google Scholar
Barton, HA, Northup, DE. Geomicrobiology in cave environments: past, current and future perspectives. Journal of Cave and Karst Studies. 2007;69(1):163–178.Google Scholar
Suzuki, S, Ishii, S, Wu, A, Cheung, A, Tenney, A, Wanger, G, et al. Microbial diversity in The Cedars, an ultrabasic, ultrareducing, and low salinity serpentinizing ecosystem. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(38):15336–15341.Google Scholar
Magnabosco, C, Tekere, M, Lau, MCY, Linage, B, Kuloyo, O, Erasmus, M, et al. Comparisons of the composition and biogeographic distribution of the bacterial communities occupying South African thermal springs with those inhabiting deep subsurface fracture water. Frontiers in Microbiology. 2014;5:679.Google Scholar
Baker, MA, Valett, HM, Dahm, CN. Organic carbon supply and metabolism in a shallow groundwater ecosystem. Ecology. 2000;81(11):3133–3148.Google Scholar
Kotelnikova, S. Microbial production and oxidation of methane in deep subsurface. Earth-Science Reviews. 2002;58(3–4):367–395.Google Scholar
Fry, NK, Fredrickson, JK, Fishbain, S, Wagner, M, Stahl, DA. Population structure of microbial communities associated with two deep, anaerobic, alkaline aquifers. Applied and Environmental Microbiology. 1997;63(4):1498–1504.Google Scholar
Vieth, A, Mangelsdorf, K, Sykes, R, Horsfield, B. Water extraction of coals – potential for estimating low molecular weight organic acids as carbon feedstock for the deep terrestrial biosphere. Organic Geochemistry. 2008;39(8):985–991.Google Scholar
Simkus, DN, Slater, GF, Lollar, BS, Wilkie, K, Kieft, TL, Magnabosco, C, et al. Variations in microbial carbon sources and cycling in the deep continental subsurface. Geochimica et Cosmochimica Acta. 2016;173:264–283.Google Scholar
Lau, MCY, Kieft, TL, Kuloyo, O, Linage-Alvarez, B, Van Heerden, E, Lindsay, MR, et al. An oligotrophic deep-subsurface community dependent on syntrophy is dominated by sulfur-driven autotrophic denitrifiers. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(49):E7927–E7936.Google Scholar
Kieft, TL, Walters, CC, Higgins, MB, Mennito, AS, Clewett, CFM, Heuer, V, et al. Dissolved organic matter compositions in 0.6–3.4 km deep fracture waters, Kaapvaal Craton, South Africa. Organic Geochemistry. 2018;118:116131.Google Scholar
Stevens, T. Lithoautotrophy in the subsurface. FEMS Microbiology Reviews. 1997;20(3–4):327–337.Google Scholar
Longnecker, K, Kujawinski, EB. Composition of dissolved organic matter in groundwater. Geochimica et Cosmochimica Acta. 2011;75(10):2752–2761.Google Scholar
Chapelle, FH, O’Neill, K, Bradley, PM, Methe, BA, Ciufo, SA, Knobel, LL, et al. A hydrogen-based subsurface microbial community dominated by methanogens. Nature. 2002;415(6869):312–315.Google Scholar
Griebler, C, Lueders, T. Microbial biodiversity in groundwater ecosystems. Freshwater Biology. 2009;54(4):649–677.Google Scholar
Flynn, TM, Sanford, RA, Ryu, H, Bethke, CM, Levine, AD, Ashbolt, NJ, et al. Functional microbial diversity explains groundwater chemistry in a pristine aquifer. BMC Microbiology. 2013;13:146.Google Scholar
Probst, AJ, Ladd, B, Jarett, JK, Geller-McGrath, DE, Sieber, CMK, Emerson, JB, et al. Differential depth distribution of microbial function and putative symbionts through sediment-hosted aquifers in the deep terrestrial subsurface. Nature Microbiology. 2018;3(3):328–336.Google Scholar
Leenheer, JA, Malcolm, RL, McKinley, PW. Occurrence of dissolved organic carbon in selected ground-water samples in the United States. Journal of Research of the US Geological Survey. 1974;2:361–369.Google Scholar
McMahon, PB, Chapelle, FH. Microbial production of organic acids in aquitard sediments and its role in aquifer geochemistry. Nature. 1991;349(6306):233–235.Google Scholar
Kusel, K, Totsche, KU, Trumbore, SE, Lehmann, R, Steinhauser, C, Herrmann, M. How deep can surface signals be traced in the critical zone? Merging biodiversity with biogeochemistry research in a central German Muschelkalk landscape. Frontiers in Earth Science. 2016;4:32.Google Scholar
Head, IM, Jones, DM, Larter, SR. Biological activity in the deep subsurface and the origin of heavy oil. Nature. 2003;426(6964):344–352.Google Scholar
Nazina, TN, Shestakova, NM, Ivoilov, VS, Kostrukov, NK, Belyaev, SS, Ivanov, MV. Radiotracer assay of microbial processes in petroleum reservoirs. Advances in Biotechnology & Microbiology. 2017;2(4):19.Google Scholar
Magot, M, Ollivier, B, Patel, BKC. Microbiology of petroleum reservoirs. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology. 2000;77(2):103–116.Google Scholar
Larter, S, Huan, H, Adams, J, Bennett, B, Jokanola, O, Oldenburg, T, et al. The controls on the composition of biodegraded oils in the deep subsurface: part II – geological controls on subsurface biodegradation fluxes and constraints on reservoir-fluid property prediction. AAPG Bulletin. 2006;90(6):921–938.Google Scholar
Zhu, YL, Vieth-Hillebrand, A, Wilke, FDH, Horsfield, B. Characterization of water-soluble organic compounds released from black shales and coals. International Journal of Coal Geology. 2015;150:265–275.Google Scholar
Onstott, TC, Lin, LH, Davidson, M, Mislowack, B, Borcsik, M, Hall, J, et al. The origin and age of biogeochemical trends in deep fracture water of the Witwatersrand Basin, South Africa. Geomicrobiology Journal. 2006;23(6):369414.Google Scholar
Kietavainen, R, Ahonen, L, Niinikoski, P, Nykanen, H, Kukkonen, IT. Abiotic and biotic controls on methane formation down to 2.5 km depth within the Precambrian Fennoscandian Shield. Geochimica et Cosmochimica Acta. 2017;202:124–145.Google Scholar
Lollar, BS, Voglesonger, K, Lin, LH, Lacrampe-Couloume, G, Telling, J, Abrajano, TA, et al. Hydrogeologic controls on episodic H2 release from Precambrian fractured rocks – energy for deep subsurface life on Earth and Mars. Astrobiology. 2007;7(6):971–986.Google Scholar
Suzuki, Y, Konno, U, Fukuda, A, Komatsu, DD, Hirota, A, Watanabe, K, et al. Biogeochemical signals from deep microbial life in terrestrial crust. PLoS One. 2014;9(12):e113063.Google Scholar
Suzuki, S, Kuenen, JG, Schipper, K, van der Velde, S, Ishii, S, Wu, A, et al. Physiological and genomic features of highly alkaliphilic hydrogen-utilizing Betaproteobacteria from a continental serpentinizing site. Nature Communications. 2014;5:3900.Google Scholar
Li, L, Wing, BA, Bui, TH, McDermott, JM, Slater, GF, Wei, S, et al. Sulfur mass-independent fractionation in subsurface fracture waters indicates a long-standing sulfur cycle in Precambrian rocks. Nature Communications. 2016;7:13252.Google Scholar
Momper, L, Jungbluth, SP, Lee, MD, Amend, JP. Energy and carbon metabolisms in a deep terrestrial subsurface fluid microbial community. ISME Journal. 2017;11(10):2319–2333.Google Scholar
Magnabosco, C, Ryan, K, Lau, MCY, Kuloyo, O, Lollar, BS, Kieft, TL, et al. A metagenomic window into carbon metabolism at 3 km depth in Precambrian continental crust. ISME Journal. 2016;10(3):730–741.Google Scholar
Colman, DR, Poudel, S, Stamps, BW, Boyd, ES, Spear, JR. The deep, hot biosphere: twenty-five years of retrospection. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(27):6895–6903.Google Scholar
Jungbluth, SP, del Rio, TG, Tringe, SG, Stepanauskas, R, Rappe, MS. Genomic comparisons of a bacterial lineage that inhabits both marine and terrestrial deep subsurface systems. PeerJ. 2017;5:e3134.Google Scholar
Lollar, BS, Lacrampe-Couloume, G, Slater, GF, Ward, J, Moser, DP, Gihring, TM, et al. Unravelling abiogenic and biogenic sources of methane in the Earth’s deep subsurface. Chemical Geology. 2006;226(3–4):328–339.Google Scholar
Kadnikov, VV, Frank, YA, Mardanov, AV, Beletskii, AV, Ivasenko, DA, Pimenov, NV, et al. Uncultured bacteria and methanogenic archaea predominate in the microbial community of western Siberian deep subsurface aquifer. Microbiology. 2017;86(3):412–415.Google Scholar
Momper, L, Reese, BK, Zinke, L, Wanger, G, Osburn, MR, Moser, D, et al. Major phylum-level differences between porefluid and host rock bacterial communities in the terrestrial deep subsurface. Environmental Microbiology Reports. 2017;9(5):501–511.Google Scholar
Wanger, G, Southam, G, Onstott, TC. Structural and chemical characterization of a natural fracture surface from 2.8 kilometers below land surface: biofilms in the deep subsurface. Geomicrobiology Journal. 2006;23(6):443–452.Google Scholar
Purkamo, L, Bomberg, M, Nyyssonen, M, Kukkonen, I, Ahonen, L, Itavaara, M. Heterotrophic communities supplied by ancient organic carbon predominate in deep fennoscandian bedrock fluids. Microbial Ecology. 2015;69(2):319–332.Google Scholar
Jangir, Y, French, S, Momper, LM, Moser, DP, Amend, JP, El-Naggar, MY. Isolation and characterization of electrochemically active subsurface Delftia and Azonexus species. Frontiers in Microbiology. 2016;7:756.Google Scholar
Lehman, RM, Colwell, FS, Bala, GA. Attached and unattached microbial communities in a simulated basalt aquifer under fracture- and porous-flow conditions. Applied and Environmental Microbiology. 2001;67(6):2799–2809.Google Scholar
Petsch, ST, Eglinton, TI, Edwards, KJ. 14C-dead living biomass: evidence for microbial assimilation of ancient organic carbon during share weathering. Science. 2001;292(5519):1127–1131.Google Scholar
Figure 0

Figure 16.1 Deep biosphere locations on the continents and in the ocean.

Figure 1

Figure 16.2 Anaerobic breakdown of OM by microorganisms via (a) methanogenesis and (b) sulfidogenesis.

Adapted from (25).
Figure 2

Figure 16.3 (a) Scanning transmission X-ray microscope image and (b) optical density map of the organic carbon distribution of sediments from 1.75 m below seafloor at Integrated Ocean Drilling Program Site 1231 Hole B, Peru Basin. The optical density map was generated by subtracting a pre-edge X-ray image from a post-edge X-ray image; brighter pixels correspond to higher concentrations of organic carbon. OM associated with particles is not distributed evenly over the surface.

Image courtesy of Dr. E. Estes, University of Delaware.
Figure 3

Figure 16.4 The abundance and composition of organic molecules in hydrothermal fluids will reflect a complex reaction history. While chemoautotrophy and abiotic synthesis involve the reduction of inorganic carbon into organic molecules, remineralization will do the reverse. Oxidation and dehydration reactions produce smaller, more polar compounds that are generally more labile and more easily consumed by heterotrophic microorganisms. Reduction and dehydration reactions may produce larger and more apolar material that is more resistant to microbial degradation and may be sequestered in the subsurface or persist for long periods of time in the deep ocean.

Figure 4

Figure 16.5 Range of methane and CO2 concentrations in basalt-hosted high-temperature (black outline; Axial Volcano, Trans-Atlantic Geotraverse (TAG), 9°N East Pacific Rise, Lucky Strike), ultramafic-hosted (green diamonds; Lost City, East Summit of Von Damm, Rainbow), ridge flank (blue checkers; Juan de Fuca ridge flank, North Pond), back-arc basins (orange diagonal; PACMANUS, Mariana Arc, Okinawa Trough), and sedimented systems (gray boxes; Guaymas, Middle Valley, Okinawa Trough). Seawater composition is included for comparison. Methane concentrations at North Pond are plotted at the reported detection limit of the analysis (0.5 µM).

References are given in Table 16.1.
Figure 5

Table 16.1 Characteristics and carbon contents of representative oceanic sub-seafloor fluids.

Figure 6

Table 16.2 Summary of characteristics and carbon contents of different types of continental subsurface systems.

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×