Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-23T13:12:35.135Z Has data issue: false hasContentIssue false

18 - The Genetics, Biochemistry, and Biophysics of Carbon Cycling by Deep Life

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

Life in the terrestrial and marine subsurface has adapted and evolved mechanisms to survive under extremes of energy limitation, temperature, pressure, radiation, and/or water availability. New developments in nucleic acid sequencing, high-pressure biochemistry, and high-pressure biophysics have expanded our understanding of the mechanisms used by deep life. This chapter synthesizes these new developments and highlights remaining gaps in understanding.

Type
Chapter
Information
Deep Carbon
Past to Present
, pp. 556 - 584
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/

18.1 Introduction

Much of the microbial life on Earth resides below the surface in the crust (Figure 18.1) (Reference Kallmeyer, Pockalny, Adhikari, Smith and D’Hondt1), either buried in marine sediments (Reference Parkes, Cragg, Bale, Getliff, Goodman and Rochelle2) and petroleum deposits (Reference Head, Jones and Röling3) or entrained in aquifers within oceanic and terrestrial rocks (Figure 18.2) (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4Reference Meyer, Jaekel, Tully, Glazer, Wheat and Lin8), fluid inclusions in salt, permafrost, and ice (Reference Mackelprang, Waldrop, Deangelis, David, Chavarria and Blazewicz9Reference Parnell and McMahon11), as well as hydrothermal and geothermal fluids (Reference Huber, Mark Welch, Morrison, Huse, Neal and Butterfield12,Reference Hedlund, Murugapiran, Alba, Levy, Dodsworth and Goertz13). The study of deep subsurface life has defined our understanding of habitability and expanded our knowledge of the mechanisms that enables life to live in these environments (Reference Hoehler and Jørgensen14). While the study of deep life may seem like a philosophical exercise, understanding this enigmatic biosphere has important real-world implications for assessing the safety and feasibility of underground storage of spent nuclear fuel and other toxic compounds, sequestration of atmospheric CO2, or acquisition of fuels such as tar sands, deep subsurface coal beds, methane hydrates, or fracking (Reference Head, Jones and Röling3,Reference Kietäväinen and Purkamo5,Reference Daly, Borton, Wilkins, Hoyt, Kountz and Wolfe15).

Figure 18.1 Global distribution of marine sediment densities of microbial cells, integrated with depth

Figure 18.2 Example schematic of a deep subsurface basaltic aquifer. Such environments provide a substantially different habitat for microbial life from sediments or soils

Organisms inhabiting subsurface environments likely have been isolated from the surface world for hundreds to millions of years (Reference Wilhelms, Larter, Head, Farrimond, Di-Primio and Zwach16). Thus, their metabolic lifestyles may differ substantially from those of surface organisms. Even though subsurface environments are diverse (Chapter 16, this volume), subsurface microbes share common biological challenges such as limitations of energy, resources, and space, as well as extremes of pressure, pH, osmolarity, and temperature (Chapter 17, this volume). On the other hand, subsurface environments offer biological advantages, too: environmental stability, protection from UV irradiation, and oxygen. These unique subsurface conditions lead to communities that are often phylogenetically and functionally diverse, with extremely slow population turnover times (Reference Hoehler and Jørgensen14,Reference Braun, Mhatre, Jaussi, Røy, Kjeldsen and Seidenkrantz17,Reference Trembath-Reichert, Morono, Ijiri, Hoshino, Dawson and Inagaki18) and efficient energy metabolisms (Reference Hoehler and Jørgensen14,Reference Larowe and Amend19). Increasingly, the roles of viruses and eukaryotes, in addition to bacteria and archaea, are being recognized in the deep subsurface biosphere (Reference Takai, Inagaki, Nakagawa, Hirayama, Nunoura and Sako20Reference Engelhardt, Kallmeyer, Cypionka and Engelen25). Several barriers hamper the study of life in Earth’s crust, such as sample acquisition and the difficulty of retrieving sterile, unaltered samples that have not been contaminated by drilling fluid. However, an even bigger hurdle is the difficulty of studying the copious subsurface microbes with no cultured representatives (Reference Hedlund, Murugapiran, Alba, Levy, Dodsworth and Goertz13,Reference Lloyd, Steen, Ladau, Yin and Crosby26). Their functional potential must be pieced together from direct assessments of biomolecules or biochemical processes in natural samples. However, even subsurface microbes related to laboratory cultures with “known” functions, may not perform those functions in the natural environment. Nonetheless, culturing has been extremely important in establishing the existence of a living deep subsurface biosphere (Reference Imachi, Aoi, Tasumi, Saito, Yamanaka and Saito27Reference Dodsworth, Ong, Williams, Dohnalkova and Hedlund32). Researchers must therefore combine direct molecular biological assessments with geochemical and geophysical environmental parameters to describe how carbon is microbially transformed in subsurface environments.

18.2 Genetic Potential of Subsurface Environments

Much research addresses the question of how subsurface life differs from surface life in genetics, biochemistry, and biophysics. Culturing-based methods are still the cornerstone of microbiology because they enable physiological assessments in carefully controlled experiments. However, it is universally accepted that culturing-based methods typically recover a very small proportion of the total microbes from any environment (Reference Staley and Konopka34Reference Hugenholtz, Goebel and Pace36). Our inability to cultivate microbes is a multifaceted problem that in part has been driven by our inability to perfectly recreate the physical and chemical conditions necessary for growth, slow microbial growth rates, and the rise in popularity of molecular techniques that produce data rapidly. Thus, like the rock record, our views of microbial life and the processes they perform in situ are incomplete. While this may seem bleak, there is renewed interest in the cultivation and development of novel culturing techniques, such as high-pressure culture vessels, and this is being spurred, in part, by discoveries made through genome sequencing.

Ribosomal gene sequencing has revolutionized our understanding of the microbial evolution and diversity of microbial phyla (Reference Woese and Fox37Reference Hugenholtz, Pitulle, Hershberger and Pace39). These culture-independent methods have been a boon for the study of subsurface life, as many of the organisms in the subsurface require direct environmental sequencing because they are slow growing (Reference Lomstein, Langerhuus, Hondt, Jørgensen and Spivack40), potentially dormant (Reference Lennon and Jones41,Reference Morita42), or require specific growth conditions or partners (Reference Lau, Kieft, Kuloyo, Linage-alvarez, Heerden and Lindsay43). Single-gene amplicon sequencing (i.e. 16S rRNA or nitrate reduction genes) remains the most used method for assessing microbial communities, but it lacks the resolution to decipher the metabolic potential of the microbes observed. However, other popular sequencing-based methods can also indicate metabolism, like single-cell amplified genomes (SAGs) or metagenomics and metatranscriptomics, where total community DNA and RNA are sequenced directly from environmental samples (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Lloyd, Schreiber, Petersen, Kjeldsen, Lever and Steen44,Reference Orsi, Edgcomb, Christman and Biddle45). The Census of Deep Life (https://vamps2.mbl.edu/portals/CODL), within the Deep Carbon Observatory, has enabled DNA sequencing from many deep subsurface environments and will provide a great resource as this work continues (Reference Sheik, Reese, Twing, Sylvan, Grim and Schrenk46).

The main hurdle for the molecular-based analysis of subsurface environments is the amount of microbial biomass present in the sample, which is often low. Nucleic acid extraction from any environmental sample is a fundamental step for all molecular-based studies (Figure 18.3). Extraction methods are extremely diverse and highly dependent on the samples being processed. Recent work has highlighted that commercial DNA/RNA extraction kit reagents, typically designed for high-biomass environments, contain their own microbiome that can skew downstream amplicon data analyses (Reference Sheik, Reese, Twing, Sylvan, Grim and Schrenk46,Reference Salter, Cox, Turek, Calus, Cookson and Moffatt47). It is also important to remove reagent contamination from SAGs, metagenomes, and metatranscriptomes (Reference Kogawa, Hosokawa, Nishikawa, Mori and Takeyama48Reference Heintz-Buschart, Yusuf, Kaysen, Etheridge, Fritz and May50). Thus, great care must be taken to screen resulting reads and assembled contigs to identify and remove suspicious sequences, while also allowing that sometimes organisms that were previously thought to represent surface contamination, such as cyanobacteria, may actually be legitimate community members (Reference Puente-sánchez, Arce-rodríguez, Oggerin, Moreno-paz, Blanco and Rodríguez51).

Figure 18.3 Schematic workflow for molecular studies of subsurface ecosystems. MAG = metagenome-assembled genome; MDA = multiple displacement amplification.

Post-sequencing analysis of genomes, metagenomes, and metatranscriptomes has made significant advances that have greatly enhanced our knowledge of subsurface life (Reference Chivian, Brodie, Alm, Culley, Dehal and DeSantis52Reference Spang, Saw, Jørgensen, Zaremba-niedzwiedzka, Martijn and Lind54). The ability to identify and extract microbial genomes from metagenomes (i.e. genome binning, also referred to as metagenome-assembled genomes (MAGs)) (Reference Dick55), along with SAGs, has enabled physiological inferences to be made for uncultured subsurface organisms. Recovery of genomes from metagenomes like Candidatus Desulforudis audaxviator (Reference Chivian, Brodie, Alm, Culley, Dehal and DeSantis52) or Leptospirillium and Ferroplasma (Reference Tyson, Chapman, Hugenholtz, Allen, Ram and Richardson56) revealed that microbes living in low-diversity, subsurface environments contain a large amount of genomic diversity, strain heterogeneity, and physiological diversity. Ca. D. audaxviator and Ca. Desulfopertinax cowenii contain metabolic pathways for heterotrophic growth as well as the potential to fix carbon via the Wood–Ljungdahl (WL) pathway (Reference Jungbluth, Glavina del Rio, Tringe, Stepanauskas and Rappé57). In subsurface environments, the WL pathway has been increasingly identified in MAGs and SAGs. The presence of the WL pathway is intuitive as it requires minimal energy to fix C to biomass (Reference Schuchmann and Müller58) and is reversible to oxidize acetate. Interestingly, energetically taxing pathways such as the Calvin–Benson–Bassham cycle are also present (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Wrighton, Castelle, Varaljay, Satagopan, Brown and Wilkins59). Metagenome sequencing and subsequent genome binning have been applied to several subsurface environments, such as sediments (marine and freshwater), aquifers, rock, caves, mines, and marine basalts (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Baker, Lazar, Teske and Dick60Reference Dodsworth, Blainey, Murugapiran, Swingley, Ross and Tringe68). In these systems, chemolithoautotrophy is always counterbalanced by heterotrophs and fermentative organisms (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Crespo-Medina, Twing, Sanchez-Murillo, Brazelton, McCollom and Schrenk61). As more subsurface environments are sampled and sequenced and their genomes assembled from metagenomes, researchers have the ability to apply pangenomic and phylogenomic approaches (Reference Delmont and Eren69) to look for functional and evolutionary differences in subsurface populations. Jungbluth et al. (Reference Jungbluth, Glavina del Rio, Tringe, Stepanauskas and Rappé57) applied these techniques to show genomically that deep terrestrial subsurface Ca. D. audaxviator lacks carbon uptake genes that the marine subsurface Ca. D. cowenii has, indicating the potential for metabolic plasticity. SAGs have also been used to target and recover specific genomes of interest from the environment (Reference Labonté, Field, Lau, Chivian, Heerden and Wommack22,Reference Lloyd, Schreiber, Petersen, Kjeldsen, Lever and Steen70,Reference Fullerton and Moyer71). These targeted approaches provided insight into the carbon metabolism of two ubiquitous and enigmatic Archaea (Reference Swan, Martinez-García, Preston, Sczyrba, Woyke and Lamy72). However, as with metagenome sequencing, recovery of complete genomes can be difficult (Reference Kogawa, Hosokawa, Nishikawa, Mori and Takeyama48). Regardless of method, these tools have provided an unprecedented glimpse into the diversity – in terms of the number of both microbes and functional genes – of subsurface microbes, and they finally allow microbiologists to draw conclusions as to the functional interactions between microbes that drive system productivity (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Baker, Lazar, Teske and Dick60,Reference Anantharaman, Brown, Hug, Sharon, Castelle and Probst64,Reference Probst, Ladd, Jarett, Geller-Mcgrath, Sieber and Emerson66).

Metatranscriptomics provides a snapshot of the environmental conditions microbes sense at the time of sampling, as the turnover of RNA in the cell is rapid, being in the order of minutes for many microbes (Reference Moran, Satinsky, Gifford, Luo, Rivers and Chan73). Orsi et al. applied a metatranscriptomic-based approach to show that sub-seafloor sediment microbes express genes for carbohydrate and protein degradation pathways (Reference Orsi, Edgcomb, Christman and Biddle45). Lau et al. (Reference Lau, Kieft, Kuloyo, Linage-alvarez, Heerden and Lindsay43) were able to construct metabolic interaction maps of microbial communities in a terrestrial subsurface South African gold mine. These microbial communities appear to use a diversity of carbon fixation and decomposition pathways to cycle carbon. De novo assembled metatranscriptomes can be a useful tool for looking at broad patterns of expression in the subsurface (Reference Lau, Kieft, Kuloyo, Linage-alvarez, Heerden and Lindsay43,Reference Jewell, Karaoz, Brodie, Williams and Beller74) or the contribution of rare community members to biogeochemical cycles (Reference Baker, Sheik, Taylor, Jain, Bhasi and Cavalcoli75). However, this method is highly dependent on the annotation databases and thus could missannotate functions or be unable to give a phylogeny. Even with these caveats, metatranscriptomics coupled to either MAGs or SAGs can link potentially novel microbes to biogeochemical cycles.

Genes and gene transcripts do not always link directly to organismal activity due to post-transcriptional regulation of microbial functions (Reference Blazewicz, Barnard, Daly and Firestone76). Therefore, proteomics may be a way to probe microbial functions of the entire community more directly. This method has only recently been applied to deep subsurface samples (Reference Lau, Kieft, Kuloyo, Linage-alvarez, Heerden and Lindsay43). Lau et al. showed that a wide variety of metabolic activities co-occurred at a 1.34‑km deep fault zone, implying that the organisms acted syntrophically to support an autotrophic-based ecosystem (Reference Lau, Kieft, Kuloyo, Linage-alvarez, Heerden and Lindsay43). Proteomics and genomics also suggested metabolic interdependencies among subsurface microbes in an acetate-fed aquifer (Reference Wrighton, Castelle, Wilkins, Hug, Sharon and Thomas77). Another method that pushes closer to measuring microbial activity in situ is metabolomics, which uses mass spectrometry to identify all of the small organic molecules present in a sample. This, too, has the potential to point to metabolic pathways that are functional in deep subsurface samples, but it has only recently come into use (Reference Bird, Tague, Zinke, Schmidt, Steen and Reese78). The greatest challenge facing metabolomics is dealing with the large number of unknown structures in the data set – a problem shared with all complex microbial communities.

The rapid acceleration of data produced from genomic, transcriptomic, and soon proteomic and metabolomic research into deep subsurface environments has opened up a new dimension regarding how we learn about this difficult-to-culture biosphere. Currently, data production far outstrips the ability of a single research lab to analyze each data set to completion. This is already beginning to support a potentially rich research area of meta-analysis studies, where combining data sets from multiple studies allows for novel insights not apparent from a single data set alone (e.g. Reference Sheik, Reese, Twing, Sylvan, Grim and Schrenk46). Through further pushing of the boundaries of what can be achieved by applying biomolecular tools to environmental samples, there is no limit to what researchers can learn about this important biosphere. Such studies not only overcome some of the limitations of whole communities full of uncultured clades, but will also lead researchers to new approaches to achieve culturing of some of these recalcitrant growers.

18.3 Biogeochemistry of Deep Subsurface Life

Microbes are Earth’s most prolific chemists. They do this by transforming organic and inorganic elements and compounds using enzymatic protein complexes. Earth’s rock record provides a glimpse of some of these processes and their evolution over time, either through isotope signatures or mineral formations, with evidence for the coevolution of minerals and biology (Reference Giovannelli, Sievert, Hügler, Markert, Becher and Schweder79,Reference Moore, Jelen, Giovannelli, Raanan and Falkowski80). While the rock record provides a hint of what processes were in operation when they were deposited, interpreting these observations relies on the characterization of extant microbes.

Microbes transform the carbon landscape of Earth’s subsurface largely by “eating” or “breathing” carbon-containing compounds (Reference Nealson81). Respiring microbes use one molecule as an electron source (reductant) and another molecule as an electron sink (oxidant). Common reductants are organic molecules such as carbohydrates, amino acids, lipids, or small molecules like hydrocarbons, acetate, and formate, as well as inorganic compounds such as reduced sulfur, ammonium, transition metals, or H2. Carbon-containing oxidants include CO2 (in the case of methanogenesis), organic molecules like fumarate or oxalate, as well as inorganic compounds like O2, oxidized sulfur or nitrogen compounds, or oxidized transition metals. An alternative to respiration is fermentation, where microbes split a single molecule of intermediate oxidation state into a more oxidized one and a more reduced one. The reduced product often serves as the reductant for another organism’s respiration. Organic matter is the most common fermentative substrate, but microbes can also ferment elemental sulfur through disproportionation. In addition to respiration and fermentation, which are primarily performed to give a microbe energy, microbes can also expend energy to take up carbon compounds. Most notably, primary producers incorporate CO2 to build biomass. Therefore, every type of microbe transforms carbon, and the presence of life on Earth has had profound effects on its carbon cycle.

To describe transformations of carbon compounds in deep subsurface environments, one must determine what combination of reductants, oxidants, and/or fermentative substrates are utilized, which microbial communities use them, and how quickly they do so (Reference Hoehler and Jørgensen14). In many deep subsurface environments, highly oxidized oxidants such as oxygen and nitrate have already been consumed by shallow organisms, leaving the less powerful (per mole) oxidants such as sulfate and CO2 to dominate (Reference Froelich, Klinkhammer, Bender, Luedtke, Heath and Cullen82,Reference D’Hondt, Rutherford and Spivack83). However, exceptions occur in environments with conduits connecting to surface-derived fluids such as deep basalts or surface-connected aquifers. In areas of the ocean floor with extremely slow sedimentation rates, O2 can be detected many meters deep into marine sediments (Reference Zinke, Reese, Mcmanus, Wheat, Orcutt and Amend33,Reference Durbin and Teske84Reference Røy, Kallmeyer, Adhikari, Pockalny, Jørgensen and D’Hondt86). CO2 fixation, usually powered by chemolithoautotrophy rather than photosynthesis, is a ubiquitous and important carbon transformation process in deep subsurface environments (Reference Magnabosco, Ryan, Lau, Kuloyo, Lollar and Kieft87).

Through the application of novel biomolecular techniques that complement the DNA, RNA, and protein-based studies descried above, great strides have been made in the understanding of the biochemistry of microbes performing carbon transformation in Earth’s subsurface. These methods include enzyme assays, lipid analysis, stable isotope probing (SIP), metabolomics of small metabolic intermediates, measuring or modeling respiration rates, and heterologous expression of novel enzymes from deep subsurface organisms.

18.3.1 Microbial Metabolism in the Deep Subsurface

Measuring the rate that a microbial function occurs is essential for determining how the function contributes to the carbon cycle. Older methods to study microbial activity in the deep biosphere, such as the incorporation of radioactive thymidine into bulk samples, have largely fallen out of favor since they are insufficiently sensitive to detect microbial activities below a few meters deep into sediments (Reference Cragg, Harvey, Fry, Herbert and Parkes88). A more sensitive method for estimating in situ microbial activity is to measure the turnover rate of trace amounts of radioactive nutrients such as electron donors/acceptors. Such methods have attained widespread use in subsurface environments and have shown that SO42– and CO2 serve as important electron acceptors for deep subsurface biosphere respiration (Reference Orcutt, LaRowe, Biddle, Colwell, Glazer and Reese89). In agreement with this, DNA signatures for microbes capable of these types of respiration have been identified in many subsurface locations (e.g. Reference Purkamo, Bomberg, Kietäväinen, Salavirta and Nyyssönen90,Reference Durbin and Teske91). Although it may seem that better oxidants should correspond to faster organic matter degradation, recent work has suggested that the rate of carbon remineralization operates independently of which oxidant is available (Reference Beulig, Røy, Glombitza and Jørgensen92). Here, radiotracer-measured rates of carbon oxidation to CO2 were constant, whether sulfate or CO2 was the terminal electron acceptor.

Direct microscopic assessment is often uninformative about either taxonomy or physiology because cells tend to be small (Reference Braun, Morono, Littmann, Kuypers, Aslan and Dong93,Reference Baker, Comolli, Dick, Hauser, Hyatt and Dill94) and their morphology does not indicate their identity or environmental functions (Reference Parkes, Cragg, Bale, Getliff, Goodman and Rochelle2). In recent years, SIP coupled to fluorescence in situ hybridization (FISH) and nanoscale secondary ion mass spectrometry (NanoSIMS) has enabled the direct metabolic assessment of natural populations incubated ex situ, but still within their natural microbial community. SIP works on the principle that organisms taking up isotopically labeled substrate will retain it in their biomolecules and can later be identified. After 9 and 405 days of seafloor incubation of a sediment core with 13C-labeled glucose, Takano et al. (Reference Takano, Chikaraishi, Ogawa, Nomaki, Morono and Inagaki95) found that subsurface archaea incorporate fresh glucose in the glycerol headgroups of their membrane lipids, but rely on detrital carbon recycling for the hydrocarbon isoprenoid groups of the lipids. FISH allows the microscopic identification of microbes of a particular taxonomic group by attaching fluorescent probes to short DNA sequences that match the taxonomically informative ribosomal RNA present in cells (Reference Amann, Fuchs and Behrens96). In places with low activity, such as the deep subsurface, the dim fluorescence signals of FISH often require amplification with catalyzed reporter deposition (CARD) (Reference Pernthaler and Amann97). NanoSIMS sputters a cesium ion beam across an area of a few micrometers to detect any isotopic label taken up by the microbe during the SIP incubation. Therefore, when combined with FISH, NanoSIMS SIP gives information about what substrates were consumed by which taxonomic groups of organisms on a nearly single-cell basis.

In 2011, Morono et al. used NanoSIMS SIP with CARD-FISH to show that cells in up to 460,000‑year-old sediments, ~200 m deep into the sub-seafloor, were capable of growing on glucose, pyruvate, and amino acids (Reference Morono, Terada, Nishizawa, Ito, Hillion and Takahata98). In addition, acetate and bicarbonate were incorporated into the biomass, although they did not promote growth. However, whether the identity of the cells that grew in these experiments reflected the community present at the time of sampling was not assessed, since only a subset of the growing cells hybridized to FISH probes (Reference Morono, Terada, Nishizawa, Ito, Hillion and Takahata98). This could be the result of either mismatches to the probe DNA sequences or difficulties of hybridizing cells during CARD-FISH (Reference Lloyd, May, Kevorkian and Steen99,Reference Buongiorno, Turner, Webster, Asai, Shumaker and Roy100).

In one of the deepest sub-seafloor drilling operations to date, researchers on Integrated Ocean Drilling Program (IODP) Expedition 337 found relatively high numbers of microbial cells whose taxonomic identities resembled terrestrial communities in 2‑km-deep coal beds, suggesting that these cells had persisted through the burial of these terrestrial sediments over tens of millions of years (Reference Inagaki, Kubo, Bowles, Heuer, Ijiri and Imachi101). By combining SIP, CARD-FISH, and NanoSIMS, Trembath-Reichert et al. found that these communities incorporated methylated compounds into the biomass, sometimes with concomitant methane production (Reference Trembath-Reichert, Morono, Ijiri, Hoshino, Dawson and Inagaki102). Ijiri et al. combined clumped isotopic measurements and label turnover rates to show that deep fluids stimulate substantial methanogenesis at mud volcanoes (Reference Ijiri, Inagaki, Kubo, Adhikari, Hattori and Hoshino103).

An analogous method to SIP that allows single-cell analysis without requiring an expensive NanoSIMS instrument is called bio-orthogonal noncanonical amino acid tagging (BONCAT) (Reference Hatzenpichler, Scheller, Tavormina, Babin, Tirrell and Orphan104). In this method, natural samples are incubated with an amino acid mimic equipped with a functional moiety that can react with a fluorophore after the incubation has ended. In this way, the microbial cells that were making new proteins (and therefore incorporated the amino acid mimic) fluoresce under a microscope and can be counted by eye or physically separated in a flow cytometer for downstream biomolecular study. BONCAT was used in subsurface samples to show stimulation of amino acid incorporation in microbial aggregates incubated with methane (Reference Hatzenpichler, Connon, Goudeau, Malmstrom, Woyke and Orphan105).

Metabolism can also be controlled by entering an inactive or dormant state. Among cultured bacteria, the Firmicutes and Actinobacteria are capable of forming endospores that can withstand low nutrients and inhospitable conditions for long time periods. These properties make sporulation a potentially advantageous phenomenon for deep subsurface life; however, it is possible that common DNA analyses miss spores since they can be difficult to lyse. Therefore, Lomstein et al. measured dipicolinic acid, a key spore coat protein, in the deep subsurface of the Peru Margin to estimate that spores in this deep subsurface environment may be as numerous as nonsporulated cells and the necromass from other dead cells that may provide food for metabolically active microbes (Reference Lomstein, Langerhuus, Hondt, Jørgensen and Spivack40). However, when the decrease in cellular biomass over sediment age is instead used to estimate the microbial necromass production rates, the rate is far outstripped by organic matter buried from ancient pelagic sedimentation (Reference Bradley, Amend and Larowe106).

In marine sediments with low organic matter contents and deep terrestrial aquifers, the radiolysis of water may provide microbial food. Natural radioactive decay of Earth’s minerals can slowly split water into H2 and oxidized moieties (Reference Lin, Hall, Lippmann-pipke, Ward, Lollar and Deflaun107). Hydrogen production from Earth’s background radiation may seem inconsequential relative to the power provided by sunlight in surface environments. However, H2 production rates from water radiolysis have been shown to be sufficient to support life in the deep subsurface (Reference Chivian, Brodie, Alm, Culley, Dehal and DeSantis52,Reference Lin, Hall, Lippmann-pipke, Ward, Lollar and Deflaun107,Reference Blair, D’Hondt, Spivack and Kingsley108). The deep subsurface may have additional sources of energy such as serpentinization, which produces the microbial substrate H2 and has been shown to support many microbial communities (Reference Ménez, Pasini and Brunelli109,Reference Suzuki, Wu, Cheung, Tenney, Wanger and Kuenen110). Serpentinization also produces a large amount of alkalinity, which inhibits autotrophy by pushing carbonate equilibrium away from CO2. Therefore, unique types of methanogens have been discovered in serpentinizing systems that have novel mechanisms for ameliorating CO2 limitation (Reference Brazelton, Schrenk, Kelley and Baross111,Reference Lang, Früh-Green, Bernasconi, Brazelton, Schrenk and McGonigle112).

A feature that may be very important to subsurface ecosystems is symbiotic relationships. Historically, microbial communities from deep subsurface aquifers have been studied by examining the cells that are caught on a 1.2‑µm filter, but recent work has shown that a wide range of ultrasmall bacteria pass through these filters (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4). These cells form previously undiscovered branches on the tree of life called candidate phyla radiations (Reference Hug, Thomas, Brown, Frischkorn, Williams and Tringe113). In addition to having extremely small cell sizes (0.009 ± 0.002 µm3), they have very small genome sizes as well, and they lack many features thought to be important to free-living organisms (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4). In addition, their cell structures show that they have cellular appendages that might aid in cell-to-cell communication, similar to those of ultrasmall uncultured archaea called archaeal Richmond Mine acidophilic nanoorganisms (ARMAN) (Figure 18.4) (Reference Baker, Comolli, Dick, Hauser, Hyatt and Dill94). Therefore, these candidate phyla radiations have been proposed to be obligate symbionts of larger microbes (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4).

Figure 18.4 Cryo-electron tomographic reconstruction of ARMAN archaea (Ar), Thermoplasmatales (Tp), vacuole (Va), and virus (Vi). (a) shows the larger image, with a 300‑nm scale bar, and (b) shows the inset, with a 50‑nm scale bar.

18.3.2 Predicting Functions of Novel Genes

Many of these genes in subsurface organisms cannot be accurately annotated by their homology to characterized gene products, since most of the latter come from microbes in pure culture (Reference Rinke, Schwientek, Sczyrba, Ivanova, Anderson and Cheng114). Given that genomic data are quickly becoming the most accessible type of data to shed light on deep biosphere communities (Section 18.2), the inability to interpret the functions of much of their genomes is a major limitation. For instance, from its genome and phylogeny, Desulfurivibrio sp. appear to be anaerobic sulfate reducers. However, when they were obtained in pure culture, they did not reduce sulfate, but instead oxidized it (Reference Thorup, Schramm, Findlay, Finster and Schreiber115). A genome from an uncultured phylum contained a homolog of the methyl coenzyme M reductase gene, from which one might conclude that it performs methanogenesis or anaerobic methane oxidation. However, enrichments of this organism could not perform methane metabolisms, but instead oxidized butyrate anaerobically (Reference Laso-pérez, Wegener, Knittel, Widdel, Harding and Krukenberg116).

One method for overcoming difficulties in accurate annotations of genes in organisms that cannot undergo phenotypic ground-truthing is to amplify target genes of interest directly from natural samples or single-cell genomes, ligate them into DNA plasmid vectors, transform them into Escherichia coli, express them as proteins, purify them, solve their structures, and characterize their activities. A predicted gene annotated as a peptidase in the uncultured phylum Bathyarchaeota was heterologously expressed in E. coli and crystallized in order to determine its structure and catalytic activity (Reference Michalska, Steen, Chhor, Endres, Webber and Bird117). Although the gene was annotated as an S15 peptidase with substrate specificity for terminal proline residues, the expressed enzyme, called bathyaminopeptidase, was found to have a much higher affinity for terminal cysteine residues. This distinction illustrates the limitation of inferring enzyme function from homology to known proteins in these uncultured deep subsurface microbes. Homologs of the gene for the II/III form of Rubisco, a well-studied key gene for CO2 fixation in the Calvin–Benson–Bassham cycle, occur in the candidate phyla radiation among the bacteria in deep terrestrial aquifers. This enzyme had previously only been found in cultured archaea, so its presence in uncultured phyla of deep subsurface bacteria suggested that perhaps it had different catalytic properties in these unknown groups (Reference Wrighton, Thomas, Sharon, Miller, Castelle and VerBerkmoes4,Reference Kantor, Wrighton, Handley, Sharon, Hug and Castelle118). However, adding the Rubisco II/III gene from the candidate phyla radiation lineage to a phototroph, Rhodobacter capsulatus, which had been engineered to remove its native Rubisco gene and had therefore lost its CO2 fixation capabilities (Reference Smith and Tabita119), caused its full CO2 fixation capabilities to be revived (Reference Wrighton, Castelle, Varaljay, Satagopan, Brown and Wilkins59).

The limitation of this heterologous gene expression approach is that it has much slower throughput than genetic characterizations. However, it provides hard evidence for function that is lacking when simply inferring enzymatic functions from homology to known proteins. In addition, it presents opportunities for discovering novel functions rather than making assumptions based on previously characterized proteins.

18.3.3 Cellular Bioenergetics

Overcoming energy limitation is a major focus of life in both marine and terrestrial subsurface environments (Reference Parnell and McMahon11,Reference Hoehler and Jørgensen14). Reaction transport modeling of oxygen and nitrate concentration profiles with sediment age showed that deep subsurface communities in the South Pacific Gyre operate at extraordinarily low metabolic rates (Reference D’Hondt, Spivack, Pockalny, Ferdelman, Fischer and Kallmeyer85,Reference Røy, Kallmeyer, Adhikari, Pockalny, Jørgensen and D’Hondt86). Subsurface microbes are therefore operating in a low energy state, and the majority of these populations may be in a long-term stationary stage in which they require energy only for cell maintenance (Reference Lever, Rogers, Lloyd, Overmann, Schink and Thauer120,Reference Jørgensen and Marshall121). Determining the minimum value for cellular maintenance is difficult, but nanocalorimetry, in which the kinetics and thermodynamics of very small populations of cells can be measured (Reference Robador, Larowe, Finkel, Amend, Nealson and Booth122), and long incubations with no new substrate additions (Reference Timmers, Gieteling, Widjaja-Greefkes, Plugge, Stams and Lens123Reference Kevorkian, Bird, Shumaker and Lloyd125) hold promise for studying organisms operating at very low energies. At the very minimum, living cells must replace biomass fast enough to overcome abiotic racemization rates, which can be very slow (Reference Steen, Jorgensen and Lomstein126,Reference Onstott, Magnabosco, Aubrey, Burton, Dworkin and Elsila127). Determining true growth rates for such natural microbial populations outside of carefully controlled laboratory conditions is nearly impossible because the change in the number of cells over time represents the product of growth rate minus death rate. However, total population turnover rates can be determined (Reference Hoehler and Jørgensen14). Braun et al. used racemization rates to estimate total population turnover times of up to tens of years for marine sediment microbes in relatively organic-rich parts of the ocean (Reference Braun, Mhatre, Jaussi, Røy, Kjeldsen and Seidenkrantz17). Slow biomass turnover times of several months to over 100 years have also been measured by measuring uptake of deuterated water (Figure 18.5) (Reference Trembath-Reichert, Morono, Ijiri, Hoshino, Dawson and Inagaki102).

Figure 18.5 Generation times for microbial life in 2-km-deep sub-seafloor coal and shale beds based on uptake rate of either 2H-labeled water or 15N-labeled compounds as described in the graph’s key

One challenge to such research is the difficulty in making accurate absolute quantifications of specific microbial populations, not just total cell counts, in natural samples that can have very low biomasses or features that disrupt quantification methods. For instance, marine sediments themselves can obscure quantifications with FISH and quantitative polymerase chain reaction (Reference Lloyd, May, Kevorkian and Steen99), and extremely small cells might be missed during filtration of water samples (Reference Baker, Comolli, Dick, Hauser, Hyatt and Dill94). However, a method that accurately quantifies the absolute cellular abundance of particular taxonomic groups in natural deep subsurface samples has yet to be developed.

18.4 Pressure Effects

All lifeforms depend on the chemistry and physics of self-replicating and self-assembling macromolecules, such as proteins, to catalyze the chemical reactions described above. The goal of research in extreme biophysics is to understand how extreme physical conditions affect the fundamental processes responsible for sustaining life. This includes understanding mechanisms of molecular adaptation to extreme conditions and the driving forces of evolution throughout geologic ages.

Much of life on Earth exists under what are considered extreme conditions of pressure: 88% of the volume of the oceans is at an average pressure of 38 MPa, the deepest trenches can exceed 100 MPa, and microbes exist at pressures as much as threefold higher deep in Earth’s crust (Reference Daniel, Oger and Winter128). The biomass in high-pressure environments may exceed that on the surface (Reference Daniel, Oger and Winter128). In the oceans, temperature decreases with depth until it reaches an almost constant 3°C with pressure as high as 1000 atmospheres (Reference Daniel, Oger and Winter128). Most of the high-pressure ocean is cold and dark, with minimal sources of carbon and minerals and with a low but constant concentration of oxygen. In contrast, temperatures of the water exiting hydrothermal vents in the deep oceans can reach 400°C, with a temperature gradient so large that the temperature decreases to 3°C only 1 m from the vent. This large temperature gradient on the ocean’s floor may be partly responsible for the highly diversified and dynamic deep subsurface microbiome. In the deep subcontinental crust, the total volume in the hydrated fissures in the rock is a large proportion of the biosphere, estimated to be 1016 m3 and 23–31 Pg of carbon (Reference Daniel, Oger and Winter128,Reference Magnabosco, Lin, Dong, Bomberg, Ghiorse and Stan-Lotter129). If only 1% of that volume is occupied by microorganisms, their biological productivity could be greater than that of Earth’s surface (Reference Daniel, Oger and Winter128). It is truly remarkable that life manages to exist in such extreme environments compared to the benign conditions in which humans exist.

18.4.1 Extreme Molecular Biophysics

The structures, interactions, and functions of biological macromolecules are highly sensitive to temperature, pressure, pH, salinity, and the concentration of small osmolytes. Take, for example, proton/sodium-coupled electron transfer (PCET) processes, which are central to energy production. The molecular machinery for H+-coupled ATP synthesis is essentially the same across most living cells; ATP synthesis succeeds regardless of physical conditions, even though the chemistry of H+-driven processes in biological macromolecules is highly sensitive to pressure, temperature, and salinity. The adaptive mechanisms for PCET under extreme conditions are unknown. Likewise, the physical properties of both the proteins and nucleic acids that drive the central dogma are highly sensitive to temperature, pressure, pH, and salinity, and yet DNA transcription and translation succeed even in extreme environments. The importance of understanding the molecular mechanisms of adaptation of fundamental biochemical processes to extreme conditions cannot be overstated. Somehow, in these extreme environments, adaptations take place so that the essential processes necessary to sustain life can proceed unimpeded.

Whereas structure–function–thermodynamics relationships in proteins from thermophiles have been studied extensively, studies on the adaptation to high pressure are limited (Reference Cavicchioli, Thomas and Curmi130). Notable exceptions are pressure-adapted RNA polymerases (Reference Kawano, Nakasone, Matsumoto, Yoshida, Usami and Kato131), metabolic enzymes (Reference Kasahara, Sato, Tamegai and Kato132Reference De Vos, Xu, Hulpiau, Vergauwen and Van Beeumen138), the bacterial chromosome organizer, histone-like nucleoid structuring proteins (Reference Jian, Xu, Gai, Xu and Xiao139), G-protein-coupled receptors in deep-sea fish (Reference Siebenaller140,Reference Siebenaller and Garrett141), and single-stranded DNA binding proteins (Reference Chilukuri, Bartlett and Fortes142). In addition to identifiable adaptations at the level of the protein’s primary sequence, pressure adaptation at the cellular level has been shown to involve changes in gene expression (Reference Boonyaratanakornkit, Miao and Clark143), such as upregulation of specific transport proteins of the outer membrane protein (OMP) family (Reference Simonato, Campanaro, Lauro, Vezzi, D’Angelo and Vitulo144,Reference Campanaro, De Pascale, Telatin, Schiavon, Bartlett and Valle145) and the production of osmolytes (Reference Samerotte, Drazen, Brand, Seibel and Yancey146Reference Bartlett149), chaperones (Reference Campanaro, Treu and Valle150,Reference Cario, Jebbar, Thiel, Kervarec and Oger151), polyamines (Reference Nakashima, Yamagami, Tomikawa, Ochi, Moriya and Asahara152), and modified lipids (Reference Bartlett149,Reference Campanaro, Treu and Valle150,Reference Michoud and Jebbar153Reference Jebbar, Franzetti, Girard and Oger155) (Figure 18.6).

Figure 18.6 Mechanisms of pressure adaptation.

Biophysical aspects of adaptation to extreme environments are poorly understood, although the effects of extreme conditions on mesophilic organisms have been studied extensively. Pressure can modulate the conformation of lipid bilayers, nucleic acids, and proteins (Reference Winter and Dzwolak156). In the case of nucleic acids, structures can be either stabilized or destabilized by pressure (Reference Macgregor157), although in general, RNA and DNA molecules exhibiting tertiary structures, such as hairpins, quadruplexes, and ribozymes, are disrupted by pressure (Reference Fan, Shek, Amiri, Dubins, Heerklotz and Macgregor158,Reference Amiri and Macgregor159). Lipid bilayers are compressed and thickened in response to pressure and become much less dynamic. The melting temperature for gel to liquid crystalline transitions increases with pressure, and organisms living at high pressures are known to have a higher proportion of unsaturated lipids in their bilayers to maintain an appropriate degree of fluidity.

It has long been known that hydrostatic pressure leads to protein unfolding (Reference Bridgman160) and oligomer dissociation (Reference Weber161,Reference Schmid, Ludemann and Jaenicke162). Pressure effects on protein structure arise from differences in molar volume between different conformational states. The molar volume of the unfolded states of proteins (U) is generally smaller than that of their folded states (F) (Figure 18.7), as is the molar volume of monomeric subunits compared to oligomeric species. Pressure shifts the folding/association equilibria toward the state that occupies the lowest molar volume: the unfolded or dissociated states. The values of these volume changes are quite small, typically <150 mL/mol for the unfolding of small globular proteins and <300 mL/mol for the dissociation of oligomers.

Figure 18.7 Schematic diagrams of a folded (a) and unfolded protein (b), in this case the pp32 leucine-rich repeat protein

(adapted from Reference Fossat, Dao, Jenkins, Dellarole, Yang and McCallum163). Cavities in (a) were calculated with a 1.4‑Å sphere using HOLLOW (Reference Ho and Gruswitz164).

Although the differences in volume between conformational or oligomeric states of proteins have been well known for some time (Reference Royer165,Reference Smeller, Biology and Smeller166), their structural origins remained obscure and were hotly debated until recently. It has been shown repeatedly that substitution of a large hydrophobic amino acid side chain such as leucine, isoleucine, or even valine by a smaller one such as alanine or glycine invariably leads to a larger overall magnitude of the volume change of unfolding and to greater sensitivity to pressure (Reference Font, Benito, Lange, Ribo and Vilanova167Reference Ando, Barstow, Baase, Fields, Matthews and Gruner169). A very strong correlation exists between the volume change of unfolding and the packing density (fractional void volume) of a given protein.

The destabilization of any given protein by pressure is larger at low temperatures. Chen and Makhatadze proposed that hydration not only does not cause pressure-induced protein unfolding, but actually makes a strong contribution that opposes pressure-induced unfolding (Reference Chen and Makhatadze170). However, as temperature increases, the increase in internal void volume becomes smaller than the effects of hydration, and thus the compensation effect diminishes and the overall difference in molar volume between folded and unfolded states diminishes (Reference Chen and Makhatadze171), and can even change sign.

Many aspects of the thermodynamic and structural pressure effects on proteins and adaptations to high-pressure environments remain unknown. It was recently shown that there is no pressure-specific “piezolytes” (i.e. there are no specific volumetric effects associated with low-molecular-weight osmolytes). Rather, their effect is simply to stabilize folded proteins against all manner of perturbations, not specifically high hydrostatic pressure (Reference Papini, Pandharipande, Royer and Makhatadze172). It is not well understood whether there are inherent, specific physical attributes of proteins and nucleic acids that allow them to tolerate high-pressure environments. On a larger scale, the ensemble of different organisms in a given environment may modulate biomolecular adaptation through the physical properties of the biomolecules. Large-scale structural adaptive differences may even be identifiable through comparison of whole genomes.

The populations of folding intermediates and low-lying alternative states implicated in the function, degradation, or aggregation of proteins from piezophiles may be different from those of their mesophilic homologs (Reference Akasaka173,Reference Akasaka, Kitahara and Kamatari174), Thus, proteins from piezophiles may have evolved to modulate the population of such species.

Thankfully, through sustained progress in instrumentation for fundamental physicochemical studies of the effects of pressure on biological molecules, the biophysics community is well positioned to address outstanding issues in biomolecular adaptation. High-pressure nuclear magnetic resonance (NMR) measurements, once restricted to a few centers with specialized equipment (Reference Jonas175Reference Kremer, Arnold, Brunner, Schuler, Jaenicke, Kalbitzer and Winter178), have been adapted for low cost and ease of use with high-field NMR spectrometers (Reference Flynn, Milton, Babu and Wand179,Reference Peterson, Nucci and Wand180). High-pressure small-angle X-ray and neutron scattering devices are available at several facilities in Japan (Reference Fujisawa, Kato and Inoko181), Europe (Reference Krywka, Sternemann, Paulus, Javid, Winter and Al-Sawalmih182), and in the United States (Reference Ando, Barstow, Baase, Fields, Matthews and Gruner169). Fluorescence, Fourier-transform infrared spectroscopy, and ultraviolet–visible spectroscopy have long been used in high-pressure biophysical research, and electron paramagnetic resonance and circular dichroism spectroscopy have recently been adapted (Reference McCoy and Hubbell183,Reference Lerch, Horwitz, McCoy and Hubbell184). Similarly, high-precision modern microcalorimeters are commercially available, as are electronically controlled, temperature-regulated cells for optical spectroscopic equipment. The range of temperature–pressure studies of biological macromolecules that are possible is broad.

18.4.2 Extreme Cellular Biophysics

In the deep-sea bacterium Photobacterium profundum (Reference Campanaro, De Pascale, Telatin, Schiavon, Bartlett and Valle145), pressure affected the expression of many genes that code for proteins implicated in nutrient transport, such as outer membrane proteins (Omps). In the yeast Saccharomyces cerevisiae (Reference Iwahashi, Odani, Ishidou and Kitagawa185), pressure affected the expression of heat-shock and metabolic proteins. Upregulation of heat-shock and cold-shock protein expression in E. coli in response to high pressure has also been reported (Reference Welch, Farewell, Neidhardt and Bartlett186). It has also been shown that pressure shock of around 1000 bar leads to the filamentation of E. coli (Reference Aertsen, Van Houdt, Vanoirbeek and Michiels187), and 40 MPa has the same effect on piezophilic Desulfovibrio indonesiensis (Reference Fichtel, Logemann, Fichtel, Rulkotter, Cypionka and Engelen188), although there are likely multiple mechanisms by which this occurs. Pressure may lead to the dissociation of FtsZ, a tubulin-type GTPase involved in forming the bacterial septum for cell division (Reference Ishii, Sato, Wachi, Nagai and Kato189). However, certain strains of E. coli bear a pressure-activated Type IV restriction endonuclease, Mrr, which leads to a pressure-dependent SOS response and subsequent filamentation (Reference Aertsen and Michiels190,Reference Aertsen and Michiels191). The molecular mechanism of the pressure-induced SOS response was shown to be due to pressure-induced dissociation of the inactive Mrr tetramer to an active Mrr dimer that cleaves cryptic sites on the E. coli chromosome (Figure 18.8).

Figure 18.8 Average fluorescence intensity images of GFP-Mrr expressed from an aranbinose-inducible promoter (PBAD) at the natural chromosomal locus in E. coli MG1655. (Left) Cells prior to the application of pressure. Full scale is 0–1.48 photon counts per 40 µs. (Right) After 10 minutes at 1 kbar and release of pressure. Full scale is 0.5–6.7 photon counts per 40 µs. Both images are 20 × 20 µm.

Although genetics, genomics, and transcriptomics can provide an overview of the putative mechanisms of adaptation to life under extreme pressures, these studies provide little biophysical insight. The study of extreme conditions on living cells remains a daunting technological challenge and requires the implementation of high-pressure quantitative microscopy approaches. A few high-pressure microscope chambers have been developed, although they have not been widely used. In the 1970s, Salmon and coworkers developed a phase contrast microscope that could withstand about 800 bar pressure (Reference Salmon and Ellis193,Reference Crenshaw and Salmon194). Müller and Gratton pioneered the use of capillary tubing for high-pressure microscopy (Reference Müller and Gratton195) that can withstand pressures up to 6000 bar. However, it has never been applied to live cell imaging. Quantitative analysis of pressure effects on bacterial mobility and flagellar rotation speed were accomplished using a recently designed high-pressure microscopy chamber (Reference Nishiyama and Kojima196). Unfortunately, owing to the thick windows of this cell, it is not suited for the high-magnification objectives required for advanced quantitative microscopy. This limitation is also the case for two other reported high-pressure microscope chamber designs (Reference Koo and Czeslik197,Reference Frey, Hartmann, Herrmann, Meyer-Pittroff, Sommer and Bluemelhuber198). Another hurdle will be the genetic manipulation of extremophile organisms in order to insert fluorescent proteins or other tags into the natural loci of genes of interest.

As in the case of high-pressure molecular biophysics, many unknowns exist regarding the molecular strategies used by cells to adapt to high-pressure or to other extreme conditions. There may be common trends (e.g. dependence on osmolytes) among different types of extremophiles. It is possible that the mechanical properties of high-pressure organisms may be different from those of their mesophilic counterparts or of hyperthermophiles, especially in their membranes and signal transduction pathways. Extreme conditions may affect cell state transitions, growth, differentiation, and development in extremophiles, or even the appearance of multicellularity. Changing geochemical conditions on Earth may have affected cellular phenotypes and the evolution of biological macromolecules.

18.5 Limits to Knowledge and Unknowns

Some of the most fundamental questions about how life on Earth evolved, especially during the first 4 billion years after Earth was formed, relate to how cells and the molecules that keep them alive adapted to changing extreme conditions. Owing to progress in genomics and in molecular and cellular biophysics, the stage is set for a rigorous and systematic examination of some of the most interesting unanswered questions about the evolution of biological systems.

Future work should focus on determining what sorts of physiological properties can be assigned to uncultured subsurface microbes, and perhaps using those insights to develop new techniques to bring them into culture. Linking the experimental work in extreme biophysics with the new frontiers being opened by coupling genomics, transcriptomics, proteomics, and metabolomics with biogeochemical analyses will yield many exiting discoveries about the deep subsurface biosphere in the next few years.

Acknowledgments

Funding provided by the Deep Carbon Observatory, especially within the Deep Life community and Census of Deep Life, has enriched much of the research described herein. Funding for this chapter was provided by an Alfred P. Sloan Research Fellowship (FG-2015-65399) and NASA Exobiology grant NNX16AL59G to KGL, NSF grants 1745061 and 1660873 to CSS, NSF MCB grant 1514575 to CAR, and NSF MCB grant 1517378 to BG-M.

Questions for the Classroom

1 What novel techniques have enabled the study of uncultured microbes in the deep subsurface?

2 What are the limitations of some of these novel techniques?

3 What are some general features of metabolism in the deep subsurface biosphere?

4 What are some key ways that high pressure impacts biomolecules?

5 What are some of the novel techniques in use for high-pressure biophysics?

References

Kallmeyer, J, Pockalny, R, Adhikari, RR, Smith, DC, D’Hondt, S. Global distribution of microbial abundance and biomass in subseafloor sediment. Proc Natl Acad Sci. 2012;109(40):16213–16216.Google Scholar
Parkes, RJ, Cragg, BA, Bale, SJ, Getliff, JM, Goodman, K, Rochelle, PA, et al. Deep bacterial biosphere in Pacific Ocean sediments. Nature. 1994;371:410–413.CrossRefGoogle Scholar
Head, IM, Jones, DM, Röling, WFM. Marine microoganisms make a meal of oil. Nat Rev Microbiol. 2006;4:173–182.CrossRefGoogle Scholar
Wrighton, KC, Thomas, BC, Sharon, I, Miller, CS, Castelle, CJ, VerBerkmoes, NC, et al. Fermentation, hydrogen, and sulfur metabolism in multiple uncultivated bacterial phyla. Science. 2012;337:1661–1665.Google Scholar
Kietäväinen, R, Purkamo, L. The origin, source, and cycling of methane in deep crystalline rock. Front Microbiol. 2015;6:725.Google Scholar
Jørgensen, SL, Zhao, R. Microbial inventory of deeply buried oceanic crust from a young ridge flank. Front Microbiol. 2016;7:820.CrossRefGoogle 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;453:59.Google Scholar
Meyer, JL, Jaekel, U, Tully, BJ, Glazer, BT, Wheat, CG, Lin, H, et al. A distinct and active bacterial community in cold oxygenated fluids circulating beneath the western flank of the Mid-Atlantic ridge. Sci Rep. 2016;6:22541.Google Scholar
Mackelprang, R, Waldrop, MP, Deangelis, KM, David, MM, Chavarria, KL, Blazewicz, SJ, et al. Metagenomic analysis of a permafrost microbial community reveals a rapid response to thaw. Nature. 2011;480(7377):368–371.CrossRefGoogle ScholarPubMed
Borgonie, G, Linage-Alverez, B, Ojo, A, Shivambu, S, Kuloyo, O, Cason, ED, Van Heerden, E. Deep subsurface mine stalactites trap endemic fissure fluid Archaea, Bacteria, and Nematoda possibly originating from ancient seas. Front Microbiol. 2015;6:833.CrossRefGoogle ScholarPubMed
Parnell, J, McMahon, S. Physical and chemical controls on habitats for life in the deep subsurface beneath continents and ice. Philos Trans R Soc A Math Phys Eng Sci. 2016;374(2059):20140293.Google Scholar
Huber, JA, Mark Welch, DB, Morrison, HG, Huse, SM, Neal, PR, Butterfield, DA, et al. Microbial population structures in the deep marine biosphere. Science. 2007;318:97100.Google Scholar
Hedlund, BP, Murugapiran, SK, Alba, TW, Levy, A, Dodsworth, JA, Goertz, GB, et al. Uncultivated thermophiles: current status and spotlight on “Aigarchaeota”. Curr Opin Microbiol. 2015;25:136–145.CrossRefGoogle ScholarPubMed
Hoehler, TM, Jørgensen, BB. Microbial life under extreme energy limitation. Nat Rev Microbiol. 2013;11(2):8394.Google Scholar
Daly, RA, Borton, MA, Wilkins, MJ, Hoyt, DW, Kountz, DJ, Wolfe, RA, et al. Microbial metabolisms in a 2.5-km-deep ecosystem created by hydraulic fracturing in shales. Nat Microbiol. 2016;1:16145.CrossRefGoogle Scholar
Wilhelms, A, Larter, SR, Head, IM, Farrimond, P, Di-Primio, R, Zwach, C. Biodegradation of oil in uplifted basins prevented by deep-burial sterilization. Nature. 2001;411:1034–1037.Google Scholar
Braun, S, Mhatre, SS, Jaussi, M, Røy, H, Kjeldsen, KU, Seidenkrantz, M, et al. Microbial turnover times in the deep seabed studied by amino acid racemization modelling. Sci Rep. 2017;7:5680.Google Scholar
Trembath-Reichert, E, Morono, Y, Ijiri, A, Hoshino, T, Dawson, KS, Inagaki, F, et al. Methyl-compound use and slow growth characterize microbial life in 2-km-deep subseafloor coal and shale beds. Proc Natl Acad Sci. 2017;114(44):E9206–E9215.CrossRefGoogle ScholarPubMed
Larowe, DE, Amend, JP. Power limits for microbial life. Front Microbiol. 2015;6:718.Google Scholar
Takai, K, Inagaki, F, Nakagawa, S, Hirayama, H, Nunoura, T, Sako, Y, et al. Isolation and phylogenetic diversity of members of previously uncultivated Epsilon-Proteobacteria in deep-sea hydrothermal fields. FEMS Microbiol Lett. 2003;218:167–174.Google Scholar
Edgcomb, VP, Beaudoin, D, Gast, R, Biddle, JF, Teske, A. Marine subsurface eukaryotes: the fungal majority. Environ Microbiol. 2011;13(1):172–183.Google Scholar
Labonté, JM, Field, EK, Lau, M, Chivian, D, Heerden, E Van, Wommack, KE, et al. Single cell genomics indicates horizontal gene transfer and viral infections in a deep subsurface Firmicutes population. Front Microbiol. 2015;6:349.Google Scholar
Anderson, RE, Beltrán, MT, Hallam, SJ, Baross, JA. Microbial community structure across fluid gradients in the Juan de Fuca Ridge hydrothermal system. FEMS Microbiol Ecol. 2013;83(2):324–339.CrossRefGoogle Scholar
Sohlberg, E, Bomberg, M, Miettinen, H, Nyyssönen, M, Salavirta, H, Vikman, M, et al. Revealing the unexplored fungal communities in deep groundwater of crystalline bedrock fracture zones in Olkiluoto, Finland. Front Microbiol. 2015;6:573.CrossRefGoogle ScholarPubMed
Engelhardt, T, Kallmeyer, J, Cypionka, H, Engelen, B. High virus-to-cell ratios indicate ongoing production of viruses in deep subsurface sediments. ISME J. 2014;8(7):1503–1509.Google Scholar
Lloyd, KG, Steen, AD, Ladau, J, Yin, J, Crosby, L. Phylogenetically novel uncultured microbial cells dominate Earth microbiomes. mSystems. 2018;3:10.1128/mSystems.00055-18.Google Scholar
Imachi, H, Aoi, K, Tasumi, E, Saito, Y, Yamanaka, Y, Saito, Y, et al. Cultivation of methanogenic community from subseafloor sediments using a continuous-flow bioreactor. ISME J. 2011;5(12):1913–1925.CrossRefGoogle ScholarPubMed
Parkes, RJ, Sellek, G, Webster, G, Martin, D, Anders, E, Weightman, AJ, et al. Culturable prokaryotic diversity of deep, gas hydrate sediments: first use of a continuous high-pressure, anaerobic, enrichment and isolation system for subseafloor sediments (DeepIsoBUG). Environ Microbiol. 2009;11:3140–3153.Google Scholar
Boone, DR, Liu, Y, Zhao, Z, Balkwill, DL, Drake, GR, Stevens, T, et al. Bacillus infernus sp. nov., an Fe(III)- and Mn(IV)-reducing anaerobe from the deep terrestrial subsurface. Int J Syst Bacteriol. 1995;45(3):441–448.Google Scholar
Russell, JA, León-zayas, R, Wrighton, K, Biddle, JF. Deep subsurface life from North Pond: enrichment, isolation, characterization and genomes of heterotrophic bacteria. Front Microbiol. 2016;65:678.Google Scholar
Bonis, BM, Gralnick, JA. Marinobacter subterrani, a genetically tractable neutrophilic Fe(II)-oxidizing strain isolated from the Soudan Iron Mine. Front Microbiol. 2015;6:719.Google Scholar
Dodsworth, JA, Ong, JC, Williams, AJ, Dohnalkova, AC, Hedlund, BP. Thermocrinis jamiesonii sp. nov., a thiosulfate-oxidizing, autotropic thermophile isolated from a geothermal spring. Int J Syst Evol Microbiol. 2015;65:4769–4775.Google Scholar
Zinke, LA, Reese, BK, Mcmanus, J, Wheat, CG, Orcutt, BN, Amend, JP. Sediment microbial communities influenced by cool hydrothermal fluid migration. Front Microbiol. 2018;9:1249.Google Scholar
Staley, JT, Konopka, A. Measurement of in situ activities of nonphotosynthetic microorganisms in aquatic and terrestrial habitats. Annu Rev Microbiol. 1985;39:321–346.Google Scholar
Rappé, MS, Giovannoni, SJ. The uncultured microbial majority. Annu Rev Microbiol. 2003;57:369–394.CrossRefGoogle ScholarPubMed
Hugenholtz, P, Goebel, BM, Pace, NR. Impact of culture-independent studies on the emerging phylogenetic view of bacterial diversity. J Bacteriol. 1998;180(18):4765–4774.CrossRefGoogle ScholarPubMed
Woese, CR, Fox, GE. Phylogenetic structure of the prokaryotic domain: the primary kingdoms. Proc Natl Acad Sci. 1977;74(11):5088–5090.CrossRefGoogle ScholarPubMed
Woese, CR. Bacterial evolution. Microbiol Rev. 1987;51(2):221–271.Google Scholar
Hugenholtz, P, Pitulle, C, Hershberger, KL, Pace, NR. Novel division level bacterial diversity in a Yellowstone hot spring. J Bacteriol. 1998;180(2):366–376.Google Scholar
Lomstein, BA, Langerhuus, AT, Hondt, SD, Jørgensen, BB, Spivack, AJ. Endospore abundance, microbial growth and necromass turnover in deep sub-seafloor sediment. Nature. 2012;484:101–104.CrossRefGoogle ScholarPubMed
Lennon, JT, Jones, SE. Microbial seed banks: the ecological and evolutionary implications of dormancy. Nat Rev Microbiol. 2011;9(2):119–130.Google Scholar
Morita, RY. Bioavailability of energy and its relationship to growth and starvation survival in nature. Can J Microbiol. 1988;34(4):436–441.Google Scholar
Lau, MCY, Kieft, TL, Kuloyo, O, Linage-alvarez, B, Heerden, E Van, Lindsay, MR, et al. An oligotrophic deep-subsurface community dependent on syntrophy is dominated by sulfur-driven autotrophic denitrifiers. Proc Natl Acad Sci. 2016;113(49):E7927–E7936.Google Scholar
Lloyd, KG, Schreiber, L, Petersen, DG, Kjeldsen, KU, Lever, MA, Steen, AD, et al. Predominant archaea in marine sediments detrital proteins. Nature. 2013;496:215–218.Google Scholar
Orsi, WD, Edgcomb, VP, Christman, GD, Biddle, JF. Gene expression in the deep biosphere. Nature. 2013;499:205–208.Google Scholar
Sheik, CS, Reese, BK, Twing, KI, Sylvan, JB, Grim, SL, Schrenk, MO, et al. Identification and removal of contaminant sequences from ribosomal gene databases: lessons from the Census of Deep Life. Front Microbiol. 2018;9:840.CrossRefGoogle ScholarPubMed
Salter, SJ, Cox, MJ, Turek, EM, Calus, ST, Cookson, WO, Moffatt, MF et al. Reagent and laboratory contamination can critically impact sequence-based microbiome analyses. BMC Biol. 2014;12:87.CrossRefGoogle ScholarPubMed
Kogawa, M, Hosokawa, M, Nishikawa, Y, Mori, K, Takeyama, H. Obtaining high-quality draft genomes from uncultured microbes by cleaning and co-assembly of single-cell amplified genomes. Sci Rep. 2018;8(1):2059.Google Scholar
Olm, MR, Butterfield, CN, Copeland, A, Boles, TC, Thomas, BC, Banfield, JF. The source and evolutionary history of a microbial contaminant identified through soil metagenomic analysis. MBio. 2017;8(1):e01969.Google Scholar
Heintz-Buschart, A, Yusuf, D, Kaysen, A, Etheridge, A, Fritz, JV, May, P, et al. Isolation of nucleic acids from low biomass samples: detection and removal of sRNA contaminants. bioRxiv. 2017;232975.Google Scholar
Puente-sánchez, AF, Arce-rodríguez, A, Oggerin, M, Moreno-paz, M, Blanco, Y, Rodríguez, N, et al. Viable cyanobacteria in the deep continental subsurface. Proc Natl Acad Sci. 2018;115(42):10702–10707.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
Hug, LA, Baker, BJ, Anantharaman, K, Brown, CT, Probst, AJ, Castelle, CJ, et al. A new view of the tree of life. Nat Microbiol. 2016;1:16048.Google Scholar
Spang, A, Saw, JH, Jørgensen, SL, Zaremba-niedzwiedzka, K, Martijn, J, Lind, AE, et al. Complex archaea that bridge the gap between prokaryotes and eukaryotes. Nature. 2015;521(7551):173–179.CrossRefGoogle ScholarPubMed
Dick, G. Genomic Approaches in Earth and Environmental Sciences. John Wiley & Sons, 2018.Google Scholar
Tyson, GW, Chapman, J, Hugenholtz, P, Allen, EE, Ram, RJ, Richardson, PM, et al. Community structure and metabolism through reconstruction of microbial genomes from the environment. Nature. 2004;428:3743.Google Scholar
Jungbluth, SP, Glavina del Rio, T, Tringe, SG, Stepanauskas, R, Rappé, MS. Genomic comparisons of a bacterial lineage that inhabits both marine and terrestrial deep subsurface systems. PeerJ. 2017;5:e3134.Google Scholar
Schuchmann, K, Müller, V. Energetics and application of heterotrophy in acteogenic bacteria. Appl Environ Microbiol. 2016;82(14):4056–4069.CrossRefGoogle ScholarPubMed
Wrighton, KC, Castelle, CJ, Varaljay, VA, Satagopan, S, Brown, CT, Wilkins, MJ, et al. RubisCO of a nucleoside pathway known from Archaea is found in diverse uncultivated phyla in bacteria. ISME J. 2016;10(11):2702–2714.Google Scholar
Baker, BJ, Lazar, CS, Teske, AP, Dick, GJ. Genomic resolution of linkages in carbon, nitrogen, and sulfur cycling among widespread estuary sediment bacteria. Microbiome. 2015;3(1):14.Google Scholar
Crespo-Medina, M, Twing, KI, Sanchez-Murillo, R, Brazelton, WJ, McCollom, TM, Schrenk, MO. Methane dynamics in a tropical serpentinizing environment: the Santa Elena ophiolite, Costa Rica. Front Microbiol. 2017;8:916.Google Scholar
Brazelton, WJ, Nelson, B, Schrenk, MO. Metagenomic evidence for H2 oxidation and H2 production by serpentinite-hosted subsurface microbial communities. Front Microbiol. 2012;2:268.CrossRefGoogle ScholarPubMed
Hamilton, TL, Jones, DS, Schaperdoth, I, Macalady, JL. Metagenomic insights into S(0) precipitation in a terrestrial subsurface lithoautotrophic ecosystem. Front Microbiol. 2015;5:756.CrossRefGoogle Scholar
Anantharaman, K, Brown, CT, Hug, LA, Sharon, I, Castelle, CJ, Probst, AJ, et al. Thousands of microbial genomes shed light on interconnected biogeochemical processes in an aquifer system. Nat Commun. 2016;7:13219.Google Scholar
Twing, KI, Brazelton, WJ, Kubo, MDY, Hyer, AJ, Cardace, D, Hoehler, TM, et al. Serpentinization-influenced groundwater harbors extremely low diversity microbial communities adapted to high pH. Front Microbiol. 2017;8:308.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. Nat Microbiol. 2018;3(3):328–336.Google Scholar
Evans, PN, Parks, DH, Chadwick, GL, Robbins, SJ, Orphan, VJ, Golding, SD, et al. Methane metabolism in the archaeal phylum Bathyarchaeota revealed by genome-centric metagenomics. Science. 2015;350(6259):434–438.Google Scholar
Dodsworth, JA, Blainey, PC, Murugapiran, SK, Swingley, WD, Ross, CA, Tringe, SG, et al. Single-cell and metagenomic analyses indicate a fermentative and saccharolytic lifestyle for members of the OP9 lineage. Nat Commun. 2013;4:1854.Google Scholar
Delmont, TO, Eren, AM. Linking pangenomes and metagenomes: the Prochlorococcus metapangenome. PeerJ. 2018;6:e4320.Google Scholar
Lloyd, KG, Schreiber, L, Petersen, DG, Kjeldsen, KU, Lever, MA, Steen, AD, et al. Predominant archaea in marine sediments degrade detrital proteins. Nature. 2013;496(7444):215–218.Google Scholar
Fullerton, H, Moyer, CL. Comparative single-cell genomics of Chloroflexi from the Okinawa trough deep-subsurface biosphere. Appl Environ Microbiol. 2016;82(10):3000–3008.Google Scholar
Swan, BK, Martinez-García, M, Preston, CM, Sczyrba, A, Woyke, T, Lamy, D, et al. Potential for chemolithoautotrophy among ubiquitous bacteria lineages in the dark ocean. Science. 2011;333(6047):1296–1300.Google Scholar
Moran, MA, Satinsky, B, Gifford, SM, Luo, H, Rivers, A, Chan, LK et al. Sizing up metatranscriptomics. ISME J. 2013;7(2):237–243.Google Scholar
Jewell, TNM, Karaoz, U, Brodie, EL, Williams, KH, Beller, HR. Metatranscriptomic evidence of pervasive and diverse chemolithoautotrophy relevant to C, S, N and Fe cycling in a shallow alluvial aquifer. ISME J. 2016;10(9):2106–2117.CrossRefGoogle Scholar
Baker, BJ, Sheik, CS, Taylor, CA, Jain, S, Bhasi, A, Cavalcoli, JD, et al. Community transcriptomic assembly reveals microbes that contribute to deep-sea carbon and nitrogen cycling. ISME J. 2013;7(10):1962–1973.Google Scholar
Blazewicz, SJ, Barnard, RL, Daly, RA, Firestone, MK. Evaluating rRNA as an indicator of microbial activity in environmental communities: limitations and uses. ISME J. 2013;7(11):2061–2068.Google Scholar
Wrighton, KC, Castelle, CJ, Wilkins, MJ, Hug, LA, Sharon, I, Thomas, BC, et al. Metabolic interdependencies between phylogenetically novel fermenters and respiratory organisms in an unconfined aquifer. ISME J. 2014;8(7):1452–1463.CrossRefGoogle Scholar
Bird, JT, Tague, E, Zinke, L, Schmidt, JM, Steen, AD, Reese, B, et al. Uncultured microbial phyla suggest mechanisms for multi-thousand-year subsistence in Baltic Sea sediments. mBio. 2019;10(2):e02376-18.Google Scholar
Giovannelli, D, Sievert, SM, Hügler, M, Markert, S, Becher, D, Schweder, T, et al. Insight into the evolution of microbial metabolism from the deep-branching bacterium, Thermovibrio ammonificans. Elife. 2017;6:e18990.Google Scholar
Moore, EK, Jelen, BI, Giovannelli, D, Raanan, H, Falkowski, PG. Metal availability and the expanding network of microbial metabolisms in the Archaean eon. Nat Geosci. 2017;10:629–636.Google Scholar
Nealson, KH. Sediment bacteria: who’s there, what are they doing, and what’s new? Annu Rev Earth Planet Sci. 1997;25:403–434.Google Scholar
Froelich, PN, Klinkhammer, GP, Bender, ML, Luedtke, NA, Heath, GR, Cullen, D, et al. Early oxidation of organic matter in pelagic sediments of the eastern equatorial Atlantic: suhoxic diagenesis. Geochim Cosmochim Acta. 1979;43(7):1075–1090.Google Scholar
D’Hondt, S, Rutherford, S, Spivack, AJ. Metabolic activity of subsurface life in deep-sea sediments. Science. 2002;295(5562):2067–2070.Google Scholar
Durbin, AM, Teske, A. Archaea in organic-lean and organic-rich marine subsurface sediments: an environmental gradient reflected in distinct phylogenetic lineages. Front Microbiol. 2012;3:168.CrossRefGoogle ScholarPubMed
D’Hondt, S, Spivack, AJ, Pockalny, R, Ferdelman, TG, Fischer, JP, Kallmeyer, J, et al. Subseafloor sedimentary life in the South Pacific Gyre. Proc Natl Acad Sci. 2009;106(28):11651–11656.Google Scholar
Røy, H, Kallmeyer, J, Adhikari, RR, Pockalny, R, Jørgensen, BB, D’Hondt, S. Aerobic microbial respiration in 86-million-year-old deep-sea red clay. Science. 2012;336(6083):922–925.CrossRefGoogle ScholarPubMed
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 J. 2015;10(3):730–741.Google Scholar
Cragg, B, Harvey, S, Fry, J, Herbert, R, Parkes, R. Bacterial biomass and activity in the deep sediment layers of the Japan Sea, Hole 798B. Proc Ocean Drill Progr. 1992;128:761–775.Google Scholar
Orcutt, BN, LaRowe, DE, Biddle, JF, Colwell, FS, Glazer, BT, Reese, BK, et al. Microbial activity in the marine deep biosphere: progress and prospects. Front Microbiol. 2013;4:115.Google Scholar
Purkamo, L, Bomberg, M, Kietäväinen, R, Salavirta, H, Nyyssönen, M. Microbial co-occurrence patterns in deep Precambrian bedrock fracture fluids. Biogeosciences. 2016;13:3091–3108.Google Scholar
Durbin, AM, Teske, A. Microbial diversity and stratification of South Pacific abyssal marine sediments. Environ Microbiol. 2011;13(12):3219–3234.Google Scholar
Beulig, F, Røy, H, Glombitza, C, Jørgensen, BB. Control on rate and pathway of anaerobic organic carbon degradation in the seabed. Proc Natl Acad Sci. 2018;115(2):367–372.Google Scholar
Braun, S, Morono, Y, Littmann, S, Kuypers, M, Aslan, H, Dong, M, et al. Size and carbon content of sub-seafloor microbial cells at Landsort Deep, Baltic Sea. Front Microbiol. 2016;7:1375.Google Scholar
Baker, BJ, Comolli, LR, Dick, GJ, Hauser, LJ, Hyatt, D, Dill, BD, et al. Enigmatic, ultrasmall, uncultivated Archaea. Proc Natl Acad Sci. 2010;107(19):8806–8811.Google Scholar
Takano, Y, Chikaraishi, Y, Ogawa, NO, Nomaki, H, Morono, Y, Inagaki, F, et al. Sedimentary membrane lipids recycled by deep-sea benthic archaea. Nat Geosci. 2010;3(12):858–861.Google Scholar
Amann, R, Fuchs, BM, Behrens, S. The identification of microorganisms by fluorescence in situ hybridisation. Curr Opin Biotechnol. 2001;12(3):231–236.Google Scholar
Pernthaler, A, Amann, R. Simultaneous fluorescence in situ hybridization of mRNA and rRNA in environmental bacteria. Appl Environ Microbiol. 2004;70(9):5426–5433.Google Scholar
Morono, Y, Terada, T, Nishizawa, M, Ito, M, Hillion, F, Takahata, N, et al. Carbon and nitrogen assimilation in deep subseafloor microbial cells. Proc Natl Acad Sci. 2011;108(45):18295–18300.Google Scholar
Lloyd, KG, May, MK, Kevorkian, RT, Steen, AD. Meta-analysis of quantification methods shows that archaea and bacteria have similar abundances in the subseafloor. Appl Environ Microbiol. 2013;79(24):7790–7799.Google Scholar
Buongiorno, J, Turner, S, Webster, G, Asai, M, Shumaker, AK, Roy, T, et al. Inter-laboratory quantification of Bacteria and Archaea in deeply buried sediments of the Baltic Sea (IODP Expedition 347). FEMS Microbiol Ecol. 2017;93(3):fix007.Google Scholar
Inagaki, F, Kubo, Y, Bowles, MW, Heuer, VB, Ijiri, A, Imachi, H, et al. Exploring deep microbial life in coal-bearing sediment down to ~2.5 km below the ocean floor. Science. 2015;349:420–424.Google Scholar
Trembath-Reichert, E, Morono, Y, Ijiri, A, Hoshino, T, Dawson, KS, Inagaki, F, et al. Methyl-compound use and slow growth characterize microbial life in 2-km-deep subseafloor coal and shale beds. Proc Natl Acad Sci. 2017;114(44):E9206–E9215.Google Scholar
Ijiri, A, Inagaki, F, Kubo, Y, Adhikari, RR, Hattori, S, Hoshino, T, et al. Deep-biosphere methane production stimulated by geofluids in the nankai accretionary complex. Sci Adv. 2018;4(6):eaao4631.Google Scholar
Hatzenpichler, R, Scheller, S, Tavormina, PL, Babin, BM, Tirrell, DA, Orphan, VJ. In situ visualization of newly synthesized proteins in environmental microbes using amino acid tagging and click chemistry. Environ Microbiol. 2014;16:2568–2590.Google Scholar
Hatzenpichler, R, Connon, SA, Goudeau, D, Malmstrom, RR, Woyke, T, Orphan, VJ. Visualizing in situ translational activity for identifying and sorting slow-growing archaeal–bacterial consortia. Proc Natl Acad Sci. 2016;113(28):E4069–E4078.Google Scholar
Bradley, JA, Amend, JP, Larowe, DE. Necromass as a limited source of energy for microorganisms in marine sediments. J Geophys Res Biogeosci. 2018;123(2):577–590.Google Scholar
Lin, L, Hall, J, Lippmann-pipke, J, Ward, JA, Lollar, BS, Deflaun, M, et al. Radiolytic H2 in continental crust: nuclear power for deep subsurface microbial communities. Geochem Geophys Geosyst. 2005;6(7):Q07003.Google Scholar
Blair, CC, D’Hondt, S, Spivack, AJ, Kingsley, RH. Radiolytic hydrogen and microbial respiration in subsurface sediments. Astrobiology. 2007;7(6):951–970.Google Scholar
Ménez, B, Pasini, V, Brunelli, D. Life in the hydrated suboceanic mantle. Nat Geosci. 2012;5(2):133–137.Google Scholar
Suzuki, S, Wu, A, Cheung, A, Tenney, A, Wanger, G, Kuenen, JG, et al. Microbial diversity in The Cedars, an ultrabasic, ultrareducing, and low salinity serpentinizing ecosystem. Proc Natl Acad Sci. 2013;110(38):15336–15341.Google Scholar
Brazelton, WJ, Schrenk, MO, Kelley, DS, Baross, JA. Methane- and sulfur-metabolizing microbial communities dominate the Lost City hydrothermal field ecosystem. Appl Environ Microbiol. 2006;72(9):6257–6270.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. Sci Rep. 2018;8(1):755.Google Scholar
Hug, LA, Thomas, BC, Brown, CT, Frischkorn, KR, Williams, KH, Tringe, SG, et al. Aquifer environment selects for microbial species cohorts in sediment and groundwater. ISME J. 2015;9(8):1846–1856.Google Scholar
Rinke, C, Schwientek, P, Sczyrba, A, Ivanova, NN, Anderson, IJ, Cheng, J-F, et al. Insights into the phylogeny and coding potential of microbial dark matter. Nature. 2013;499(7459):431–437.Google Scholar
Thorup, C, Schramm, A, Findlay, AJ, Finster, KW, Schreiber, L. Disguised as a sulfate reducer: growth of the deltaproteobacterium Desulfurivibrio alkaliphilus by sulfide oxidation with nitrate. MBio. 2017;8(4):e00671-17.Google Scholar
Laso-pérez, R, Wegener, G, Knittel, K, Widdel, F, Harding, KJ, Krukenberg, V, et al. Thermophilic archaea activate butane via alkyl-coenzyme M formation. Nature. 2016;539(7629):396401.Google Scholar
Michalska, K, Steen, AD, Chhor, G, Endres, M, Webber, AT, Bird, J, et al. New aminopeptidase from “microbial dark matter” archaeon. FASEB J. 2015;29:4071–4079.Google Scholar
Kantor, RS, Wrighton, KC, Handley, KM, Sharon, I, Hug, LA, Castelle, CJ, et al. Small genomes and sparse metabolisms of sediment-associated bacteria from four candidate phyla. MBio. 2013;4(5):e00708–e00713.Google Scholar
Smith, SA, Tabita, FR. Positive and negative selection of mutant forms of prokaryotic (cyanobacterial) ribulose-1,5,bisphosphate carboxylase/oxygenase. J Mol Biol. 2003;331:557–569.Google Scholar
Lever, MA, Rogers, KL, Lloyd, KG, Overmann, J, Schink, B, Thauer, RK, et al. Life under extreme energy limitation: a synthesis of laboratory- and field-based investigations. FEMS Microbiol Rev. 2015;39(5):688728.Google Scholar
Jørgensen, BB, Marshall, IPG. Slow microbial life in the seabed. Annu Rev Mar Sci. 2016;8:311–332.Google Scholar
Robador, A, Larowe, DE, Finkel, SE, Amend, JP, Nealson, KH, Booth, V. Changes in microbial energy metabolism measured by nanocalorimetry during growth phase transitions. Front Microbiol. 2018;9:109.CrossRefGoogle ScholarPubMed
Timmers, PH, Gieteling, J, Widjaja-Greefkes, HCA, Plugge, CM, Stams, AJM, Lens, PNL, et al. Growth of anaerobic methane-oxidizing archaea and sulfate-reducing bacteria in a high-pressure membrane capsule bioreactor. Appl Environ Microbiol. 2015;81(4):1286–1296.Google Scholar
Finkel, SE. Long-term survival during stationary phase: evolution and the GASP phenotype. Nat Rev Microbiol. 2006;4(2):113–120.Google Scholar
Kevorkian, R, Bird, JT, Shumaker, A, Lloyd, KG. Estimating population turnover rates by relative quantification methods reveals microbial dynamics in marine sediment. Appl Environ Microbiol. 2018;84(1):e01443-17.Google Scholar
Steen, AD, Jorgensen, BB, Lomstein, BA. Abiotic racemization kinetics of amino acids in marine sediments. PLoS One. 2013;8(8):e71648.Google Scholar
Onstott, TC, Magnabosco, C, Aubrey, AD, Burton, AS, Dworkin, JP, Elsila, JE, et al. Does aspartic acid racemization constrain the depth limit of the subsurface biosphere? Geobiology. 2014;12:119.Google Scholar
Daniel, I, Oger, P, Winter, R. Origins of life and biochemistry under high-pressure conditions. Chem Soc Rev. 2006;35(10):858–875.Google Scholar
Magnabosco, C, Lin, L-H, Dong, H, Bomberg, M, Ghiorse, W, Stan-Lotter, H, et al. The biomass and biodiversity of the continental subsurface. Nat Geosci. 2018;11:707–717.Google Scholar
Cavicchioli, R, Thomas, T, Curmi, PMG. Cold stress response in Archaea. Extremophiles. 2000;4(6):321–331.Google Scholar
Kawano, H, Nakasone, K, Matsumoto, M, Yoshida, Y, Usami, R, Kato, C, et al. Differential pressure resistance in the activity of RNA polymerase isolated from Shewanella violacea and Escherichia coli. Extremophiles. 2004;8(5):367–375.Google Scholar
Kasahara, R, Sato, T, Tamegai, H, Kato, C. Piezo-adapted 3-isopropylmalate dehydrogenase of the obligate piezophile Shewanella benthica DB21MT-2 isolated from the 11,000-m depth of the Mariana Trench. Biosci Biotechnol Biochem. 2009;73(11):2541–2543.Google Scholar
Hay, S, Evans, RM, Levy, C, Loveridge, EJ, Wang, X, Leys, D, et al. Are the catalytic properties of enzymes from piezophilic organisms pressure adapted? ChemBioChem. 2009;10(14):2348–2353.CrossRefGoogle ScholarPubMed
Nishiguchi, Y, Abe, F, Okada, M. Different pressure resistance of lactate dehydrogenases from hagfish is dependent on habitat depth and caused by tetrameric structure dissociation. Mar Biotechnol. 2011;13(2):137–141.CrossRefGoogle ScholarPubMed
Hamajima, Y, Nagae, T, Watanabe, N, Ohmae, E, Kato-Yamada, Y, Kato, C. Pressure adaptation of 3-isopropylmalate dehydrogenase from an extremely piezophilic bacterium is attributed to a single amino acid substitution. Extremophiles. 2016;20(2):177–186.Google Scholar
Rosenbaum, E, Gabel, F, Durá, MA, Finet, S, Cléry-Barraud, C, Masson, P, et al. Effects of hydrostatic pressure on the quaternary structure and enzymatic activity of a large peptidase complex from Pyrococcus horikoshii. Arch Biochem Biophys. 2012;517(2):104–110.Google Scholar
Murakami, C, Ohmae, E, Tate, S, Gekko, K, Nakasone, K, Kato, C. Comparative study on dihydrofolate reductases from Shewanella species living in deep-sea and ambient atmospheric-pressure environments. Extremophiles. 2011;15(2):165–175.Google Scholar
De Vos, D, Xu, Y, Hulpiau, P, Vergauwen, B, Van Beeumen, JJ. Structural investigation of cold activity and regulation of aspartate carbamoyltransferase from the extreme psychrophilic bacterium Moritella profunda. J Mol Biol. 2007;365(2):379–395.Google Scholar
Jian, H, Xu, G, Gai, Y, Xu, J, Xiao, X. The histone-like nucleoid structuring protein (H-NS) is a negative regulator of the lateral flagellar system in the deep-sea bacterium Shewanella piezotolerans WP3. Appl Environ Microbiol. 2016;82(8):2388–2398.Google Scholar
Siebenaller, JF. Pressure effects on the GTPase activity of brain membrane G proteins of deep-living marine fishes. Comp Biochem Physiol B Biochem Mol Biol. 2003;135(4):697705.Google Scholar
Siebenaller, JF, Garrett, DJ. The effects of the deep-sea environment on transmembrane signaling. Comp Biochem Physiol B Biochem Mol Biol. 2002;131(4):675–694.Google Scholar
Chilukuri, LN, Bartlett, DH, Fortes, PAG. Comparison of high pressure-induced dissociation of single-stranded DNA-binding protein (SSB) from high pressure-sensitive and high pressure-adapted marine Shewanella species. Extremophiles. 2002;6(5):377–383.Google Scholar
Boonyaratanakornkit, BB, Miao, LY, Clark, DS. Transcriptional responses of the deep-sea hyperthermophile Methanocaldococcus jannaschii under shifting extremes of temperature and pressure. Extremophiles. 2007;11(3):495503.Google Scholar
Simonato, F, Campanaro, S, Lauro, FM, Vezzi, A, D’Angelo, M, Vitulo, N, et al. Piezophilic adaptation: a genomic point of view. J Biotechnol. 2006;126(1):1125.Google Scholar
Campanaro, S, De Pascale, F, Telatin, A, Schiavon, R, Bartlett, DH, Valle, G. The transcriptional landscape of the deep-sea bacterium Photobacterium profundum in both a toxR mutant and its parental strain. BMC Genomics. 2012;13:567.Google Scholar
Samerotte, AL, Drazen, JC, Brand, GL, Seibel, BA, Yancey, PH. Correlation of trimethylamine oxide and habitat depth within and among species of teleost fish: an analysis of causation. Physiol Biochem Zool. 2007;80(2):197208.Google Scholar
Nagae, T, Kawamura, T, Chavas, LMG, Niwa, K, Hasegawa, M, Kato, C, et al. High-pressure-induced water penetration into 3-isopropylmalate dehydrogenase. Acta Crystallogr Sect D Biol Crystallogr. 2012;68(3):300–309.Google Scholar
Kaye, JZ, Baross, J. Synchronous effects of temperature, hydrostatic pressure, and salinity on growth, phospholipid profiles, and protein patterns of four Halomonas species isolated from deep-sea hydrothermal-vent and sea surface environments. Appl Environ Microbiol. 2004;70(10):6220–6229.Google Scholar
Bartlett, DH. Pressure effects on in vivo microbial processes. Biochim Biophys Acta Protein Struct Mol Enzymol. 2002;1595(1):367–381.Google Scholar
Campanaro, S, Treu, L, Valle, G. Protein evolution in deep sea bacteria: an analysis of amino acids substitution rates. BMC Evol Biol. 2008;8(1):313.Google Scholar
Cario, A, Jebbar, M, Thiel, A, Kervarec, N, Oger, PM. Molecular chaperone accumulation as a function of stress evidences adaptation to high hydrostatic pressure in the piezophilic archaeon Thermococcus barophilus. Sci Rep. 2016;6(1):29483.Google Scholar
Nakashima, M, Yamagami, R, Tomikawa, C, Ochi, Y, Moriya, T, Asahara, H, et al. Long and branched polyamines are required for maintenance of the ribosome, tRNAHis and tRNATyr in Thermus thermophilus cells at high temperatures. Genes Cells. 2017;22(7):628–645.Google Scholar
Michoud, G, Jebbar, M. High hydrostatic pressure adaptive strategies in an obligate piezophile Pyrococcus yayanosii. Sci Rep. 2016;6:27289.Google Scholar
Eloe, EA, Malfatti, F, Gutierrez, J, Hardy, K, Schmidt, WE, Pogliano, K, et al. Isolation and characterization of a psychropiezophilic alphaproteobacterium. Appl Environ Microbiol. 2011;77(22):8145–8153.Google Scholar
Jebbar, M, Franzetti, B, Girard, E, Oger, P. Microbial diversity and adaptation to high hydrostatic pressure in deep-sea hydrothermal vents prokaryotes. Extremophiles. 2015;19(4):721–740.Google Scholar
Winter, R, Dzwolak, W. Exploring the temperature-pressure configurational landscape of biomolecules: from lipid membranes to proteins. Philos Trans A Math Phys Eng Sci. 2005;363(1827):537–562; discussion 562–563.Google Scholar
Macgregor, RB. Effect of hydrostatic, pressure on nucleic acids. Biopolymers. 1998;48(4):253–263.Google Scholar
Fan, HY, Shek, YL, Amiri, A, Dubins, DN, Heerklotz, H, Macgregor, RB Jr., et al. Volumetric characterization of sodium-induced G-quadruplex formation. J Am Chem Soc. 2011;133(12):4518–4526.Google Scholar
Amiri, AR, Macgregor, RB. Jr. The effect of hydrostatic pressure on the thermal stability of DNA hairpins. Biophys Chem. 2011;156(1):8895.Google Scholar
Bridgman, . The coagulation of albumin by pressure. J Biol Chem. 1914;19:511–512.Google Scholar
Weber, G. Thermodynamics of the association and the pressure dissociation of oligomeric proteins. J Phys Chem. 1993;97(27):7108–7115.Google Scholar
Schmid, G, Ludemann, H-D, Jaenicke, R. Dissociationand aggregation of lactic dehydrogenase by high hydrostatic pressure. Eur J Biochem. 1979;97:407–413.Google Scholar
Fossat, MJ, Dao, TP, Jenkins, K, Dellarole, M, Yang, Y, McCallum, SA, et al. High-resolution mapping of a repeat protein folding free energy landscape. Biophys J. 2016;111(11):2368–2376.Google Scholar
Ho, BK, Gruswitz, F. HOLLOW: generating accurate representations of channel and interior surfaces in molecular structures. BMC Struct Biol. 2008;6:49.Google Scholar
Royer, CA. Revisiting volume changes in pressure-induced protein unfolding. Biochim Biophys Acta Protein Struct Mol Enzymol. 2002;1595(1–2):201–209.Google Scholar
Smeller, L, Biology, R, Smeller, L. Pressure–temperature phase diagrams of biomolecules. Biochim Biophys Acta Protein Struct Mol Enzymol. 2002;1595(1–2):1129.Google Scholar
Font, J, Benito, A, Lange, R, Ribo, M, Vilanova, M. The contribution of the residues from the main hydrophobic core of ribonuclease A to its pressure-folding transition state. Protein Sci. 2006;15(5):1000–1009.Google Scholar
Roche, J, Caro, JA, Norberto, DR, Barthe, P, Roumestand, C, Schlessman, JL, et al. Cavities determine the pressure unfolding of proteins. Proc Natl Acad Sci. 2012;109(18):6945–6950.Google Scholar
Ando, N, Barstow, B, Baase, WA, Fields, A, Matthews, BW, Gruner, SM. Structural and thermodynamic characterization of T4 lysozyme mutants and the contribution of internal cavities to pressure denaturation. Biochemistry. 2008;47(42):11097–11109.Google Scholar
Chen, CR, Makhatadze, GI. Molecular determinant of the effects of hydrostatic pressure on protein folding stability. Nat Commun. 2017;8:14561.Google Scholar
Chen, CR, Makhatadze, GI. Molecular determinants of temperature dependence of protein volume change upon unfolding. J Phys Chem B. 2017;121(35):8300–8310.Google Scholar
Papini, CM, Pandharipande, PP, Royer, CA, Makhatadze, GI. Putting the piezolyte hypothesis under pressure. Biophys J. 2017;113(5):974–977.Google Scholar
Akasaka, K. Probing conformational fluctuation of proteins by pressure perturbation. Chem Rev. 2006;106(5):1814–1835.Google Scholar
Akasaka, K, Kitahara, R, Kamatari, YO. Exploring the folding energy landscape with pressure. Arch Biochem Biophys. 2013;531(1–2):110–115.Google Scholar
Jonas, J. High-resolution nuclear magnetic resonance studies of proteins. Biochim Biophys Acta Protein Struct Mol Enzymol. 2002;1595(1–2):145–159.Google Scholar
Peng, X, Jonas, J, Silva, JL. High-pressure NMR study of the dissociation of arc repressor. Biochemistry. 1994;33(27):8323–8329.Google Scholar
Li, H, Yamada, H, Akasaka, K. Effect of pressure on individual hydrogen bonds in proteins. Basic pancreatic trypsin inhibitor. Biochemistry. 1998;37(5):1167–1173.Google Scholar
Kremer, W, Arnold, MR, Brunner, E, Schuler, B, Jaenicke, R, Kalbitzer, HR. High pressure NMR spectroscopy and its application to the cold shock protein TmCsp derived from the hyperthermophilic bacterium Thermotoga maritima. In: Winter, R, ed. Advances in High Pressure Bioscience and Biotechnology II, Proceedings. Springer, 2003, pp. 101–112.Google Scholar
Flynn, PF, Milton, MJ, Babu, CR, Wand, AJ. A simple and effective NMR cell for studies of encapsulated proteins dissolved in low viscosity solvents. J Biomol NMR. 2002;23(4):311–316.Google Scholar
Peterson, RW, Nucci, NV, Wand, AJ. Modification of encapsulation pressure of reverse micelles in liquid ethane. J Magn Reson. 2011;212(1):229–233.Google Scholar
Fujisawa, T, Kato, M, Inoko, Y. Structural characterization of lactate dehydrogenase dissociation under high pressure studied by synchrotron high-pressure small-angle X-ray scattering. Biochemistry. 1999;38(20):6411–6418.Google Scholar
Krywka, C, Sternemann, C, Paulus, M, Javid, N, Winter, R, Al-Sawalmih, A, et al. The small-angle and wide-angle X-ray scattering set-up at beamline BL9 of DELTA. J Synchrotron Radiat. 2007;14(3):244–251.Google Scholar
McCoy, J, Hubbell, WL. High-pressure EPR reveals conformational equilibria and volumetric properties of spin-labeled proteins. Proc Natl Acad Sci. 2011;108(4):1331–1336.Google Scholar
Lerch, MT, Horwitz, J, McCoy, J, Hubbell, WL. Circular dichroism and site-directed spin labeling reveal structural and dynamical features of high-pressure states of myoglobin. Proc Natl Acad Sci. 2013;110(49):E4714–E4722.Google Scholar
Iwahashi, H, Odani, M, Ishidou, E, Kitagawa, E. Adaptation of Saccharomyces cerevisiae to high hydrostatic pressure causing growth inhibition. FEBS Lett. 2005;579(13):2847–2852.Google Scholar
Welch, TJ, Farewell, A, Neidhardt, FC, Bartlett, DH. Stress response of Escherichia coli to elevated hydrostatic pressure. J Bacteriol. 1993;175(22):7170–7177.Google Scholar
Aertsen, A, Van Houdt, R, Vanoirbeek, K, Michiels, CW. An SOS response induced by high pressure in Eschericia coli. J Bacteriol. 2004;186(18):6133–6141.Google Scholar
Fichtel, K, Logemann, J, Fichtel, J, Rulkotter, J, Cypionka, H, Engelen, B. Temperature and pressure adaptation of a sulfate reducer from the deep subsurface. Front Microbiol. 2015;6:1078.Google Scholar
Ishii, A, Sato, T, Wachi, M, Nagai, K, Kato, C. Effects of high hydrostatic pressure on bacterial cytoskeleton FtsZ polymers in vivo and in vitro. Microbiology. 2004;150(Pt 6):1965–1972.Google Scholar
Aertsen, A, Michiels, CW. Mrr instigates the SOS response after high pressure stress in Escherichia coli. Mol Microbiol. 2005;58(5):1381–1391.Google Scholar
Aertsen, A, Michiels, CW. SulA-dependent hypersensitivity to high pressure and hyperfilamentation after high-pressure treatment of Escherichia coli lon mutants. Res Microbiol. 2005;156(2):233–237.Google Scholar
Bourges, AC, Torres Montaguth, OE, Ghosh, A, Tadesse, WM, Declerck, N, Aertsen, A, et al. High pressure activation of the Mrr restriction endonuclease in Escherichia coli involves tetramer dissociation. Nucleic Acids Res. 2017;45(9):5323–5332.Google Scholar
Salmon, ED, Ellis, GW. A new miniature chamber pressure for microscopy strain-free optical glass windows facilitate phase-contrast and polarized-light microscopy of living cells. J Cell Biol. 1975;65:587602.Google Scholar
Crenshaw, HC, Salmon, ED. Hydrostatic pressure to 400 atm does not induce changes in the cytosolic concentration of Ca2+ in mouse fibroblasts: measurements using fura-2 fluorescence. Exp Cell Res. 1996;227(2):277–284.Google Scholar
Müller, JD, Gratton, E. High-pressure fluorescence correlation spectroscopy. Biophys J. 2003;85(4):2711–2719.Google Scholar
Nishiyama, M, Kojima, S. Bacterial motility measured by a miniature chamber for high-pressure microscopy. Int J Mol Sci. 2012;13(7):9225–9239.Google Scholar
Koo, J, Czeslik, C. High pressure sample cell for total internal reflection fluorescence spectroscopy at pressures up to 2500 bar. Rev Sci Instrum. 2012;83(8):085109.Google Scholar
Frey, B, Hartmann, M, Herrmann, M, Meyer-Pittroff, R, Sommer, K, Bluemelhuber, G. Microscopy under pressure – an optical chamber system for fluorescence microscopic analysis of living cells under high hydrostatic pressure. Microsc Res Tech. 2006;69(2):6572.Google Scholar
Figure 0

Figure 18.1 Global distribution of marine sediment densities of microbial cells, integrated with depth

(adapted from 1).
Figure 1

Figure 18.2 Example schematic of a deep subsurface basaltic aquifer. Such environments provide a substantially different habitat for microbial life from sediments or soils

(adapted from 33).
Figure 2

Figure 18.3 Schematic workflow for molecular studies of subsurface ecosystems. MAG = metagenome-assembled genome; MDA = multiple displacement amplification.

Figure 3

Figure 18.4 Cryo-electron tomographic reconstruction of ARMAN archaea (Ar), Thermoplasmatales (Tp), vacuole (Va), and virus (Vi). (a) shows the larger image, with a 300‑nm scale bar, and (b) shows the inset, with a 50‑nm scale bar.

Image reproduced from (94).
Figure 4

Figure 18.5 Generation times for microbial life in 2-km-deep sub-seafloor coal and shale beds based on uptake rate of either 2H-labeled water or 15N-labeled compounds as described in the graph’s key

(adapted from 102).
Figure 5

Figure 18.6 Mechanisms of pressure adaptation.

Figure 6

Figure 18.7 Schematic diagrams of a folded (a) and unfolded protein (b), in this case the pp32 leucine-rich repeat protein

(adapted from 163). Cavities in (a) were calculated with a 1.4‑Å sphere using HOLLOW (164).
Figure 7

Figure 18.8 Average fluorescence intensity images of GFP-Mrr expressed from an aranbinose-inducible promoter (PBAD) at the natural chromosomal locus in E. coli MG1655. (Left) Cells prior to the application of pressure. Full scale is 0–1.48 photon counts per 40 µs. (Right) After 10 minutes at 1 kbar and release of pressure. Full scale is 0.5–6.7 photon counts per 40 µs. Both images are 20 × 20 µm.

Results are similar to those published in (192).

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
×