Zetaproteobacteria

Last updated

Zetaproteobacteria
Scientific classification
Domain:
Phylum:
Class:
Zetaproteobacteria

Makita et al. 2017
Order:
Mariprofundales

Makita et al. 2017
Family:
Mariprofundaceae

Hördt et al. 2020
Genera
Synonyms
  • "Mariprofundia" Cavalier-Smith 2020

The class Zetaproteobacteria is the sixth and most recently described class of the Pseudomonadota. [1] Zetaproteobacteria can also refer to the group of organisms assigned to this class. The Zetaproteobacteria were originally represented by a single described species, Mariprofundus ferrooxydans , [2] which is an iron-oxidizing neutrophilic chemolithoautotroph originally isolated from Kamaʻehuakanaloa Seamount (formerly Loihi) in 1996 (post-eruption). [1] [3] Molecular cloning techniques focusing on the small subunit ribosomal RNA gene have also been used to identify a more diverse majority of the Zetaproteobacteria that have as yet been unculturable. [4]

Contents

Regardless of culturing status, the Zetaproteobacteria show up worldwide in estuarine and marine habitats associated with opposing steep redox gradients of reduced (ferrous) iron and oxygen, either as a minor detectable component or as the dominant member of the microbial community. [5] [6] [7] [8] [9] [10] Zetaproteobacteria have been most commonly found at deep-sea hydrothermal vents, [4] though recent discovery of members of this class in near-shore environments has led to the reevaluation of Zetaproteobacteria distribution and significance. [11] [12] [13]

Microbial mats encrusted with iron oxide on the flank of Kama`ehuakanaloa Seamount, Hawaii. Microbial communities in this type of habitat can harbor microbial communities dominated by the iron-oxidizing Zetaproteobacteria. Loihiflank.jpg
Microbial mats encrusted with iron oxide on the flank of Kamaʻehuakanaloa Seamount, Hawaii. Microbial communities in this type of habitat can harbor microbial communities dominated by the iron-oxidizing Zetaproteobacteria.

Significance

The Zetaproteobacteria are distributed worldwide in deep sea and near shore environments at oxic/anoxic interfaces. With this wide distribution, the Zetaproteobacteria have the potential to play a substantial role in biogeochemical cycling, both past and present. Ecologically, the Zetaproteobacteria play a major role in the engineering of their own environment through the use of the controlled deposition of mineralized iron oxides, also directly affecting the environment of other members of the microbial community.

Prevalence of the Zetaproteobacteria in near-shore metal (e.g. steel) coupon biocorrosion experiments highlights the impact of these marine iron oxidizers on expensive problems such as the rusting of ship hulls, metal pilings and pipelines. [11] [14] [15]

Discovery

Mariprofundus ferrooxydans PV-1 twisted stalks TEM image. One example of Fe oxide morphotypes produced by the Zetaproteobacteria. Image by Clara Chan Mariprofundus ferrooxydans PV-1 stalks TEM image.tif
Mariprofundus ferrooxydans PV-1 twisted stalks TEM image. One example of Fe oxide morphotypes produced by the Zetaproteobacteria. Image by Clara Chan

The Zetaproteobacteria were first discovered in 1991 by Craig Moyer, Fred Dobbs and David Karl as a single rare clone in a mesophilic, or moderate temperature, hydrothermal vent field known as Pele's Vents at Kamaʻehuakanaloa Seamount (formerly Loihi), Hawaii. This particular vent was dominated by sulfur-oxidizing Campylobacterota. With no close relatives known at the time, the clone was initially labeled as Gammaproteobacteria. [16]

Subsequent isolation of two strains of M. ferrooxydans, PV-1 and JV-1, [3] along with the increasing realization that a phylogenetically distinct group of Pseudomonadota (the Zetaproteobacteria) could be found globally as dominant members of bacterial communities led to the suggestion for the creation of this new class of the Proteobacteria.

Cultivation

Neutrophilic microaerophilic Fe-oxidizing bacteria are typically cultivated using an agarose-stabilized or liquid culture with an FeS or FeCO3 plug. The headspace of the culture tube is then purged with air or a low concentration of oxygen (often 1% or less O2). Fe-oxidizers have also successfully been cultivated in liquid culture with FeCl2 as the Fe source. These cultivation techniques follow those found in Emerson and Floyd (2005). [17]

Recently, researchers have been able to culture the Zetaproteobacteria using graphite electrodes at a fixed voltage. [18] Researchers have also aimed to improve cultivation techniques using a high-biomass batch culturing technique. [19]

Morphology

One of the most distinctive ways of identifying circumneutral iron oxidizing bacteria visually is by identifying the structure of the mineralized iron oxyhydroxide product created during iron oxidation. [3] [20] Oxidized, or ferric, iron is insoluble at circumneutral pH, thus the microbe must have a way of dealing with the mineralized "waste" product. It is thought that one method to accomplish this is to control the deposition of oxidized iron. [21] [22] [23] Some of the most common morphotypes include: amorphous particulate oxides, twisted or helical stalks (figure), [21] sheaths, [24] and y-shaped irregular filaments.

These morphologies exist both in freshwater and marine iron habitats, though common freshwater iron-oxidizing bacteria such as Gallionella sp. (twisted stalk) and Leptothrix ochracea (sheath) have only extremely rarely been found in the deep sea (not significant abundance). One currently published morphotype that has been partially resolved is the twisted stalk, which is commonly formed by M. ferrooxydans. This bacteria is a gram-negative kidney-bean-shaped cell that deposits iron oxides on the concave side of the cell, forming twisted stalks as it moves through its environment. [21] [22]

Mariprofundus ferrooxydans PV-1 cell attached to twisted stalk TEM image. Image by Clara Chan. Mariprofundus ferrooxydans PV-1 cell and stalk TEM image.tiff
Mariprofundus ferrooxydans PV-1 cell attached to twisted stalk TEM image. Image by Clara Chan.

Another common Zetaproteobacteria morphotype is the sheath structure, which has yet to be isolated, but has been identified with fluorescence in situ hybridization (FISH). [24]

Iron oxidation morphotypes can be preserved and have been detected in ancient hydrothermal deposits preserved in the rock record. [25] [26] [27] [28] [29] Some current work is focused on how the Zetaproteobacteria form their individual biominerals in the modern environment so that scientists can better interpret Fe biominerals found in the rock record. [30] [31] [32]

Ecology

Phylogenetic tree showing the phylogenetic placement of the Zetaproteobacteria (orange branches) within the Pseudomonadota. Asterisks highlight the Zetaproteobacteria cultured isolates. Phylogenetic Tree focusing on Zetaproteobacteria (png smaller).png
Phylogenetic tree showing the phylogenetic placement of the Zetaproteobacteria (orange branches) within the Pseudomonadota. Asterisks highlight the Zetaproteobacteria cultured isolates.

Biodiversity

An operational taxonomic unit, or an OTU, allows a microbiologist to define a bacterial taxa using defined similarity bins based on a gene of interest. In microbial ecology, the small subunit ribosomal RNA gene is generally used at a cut off of 97% similarity to define an OTU. In the most basic sense, the OTU represents a bacterial species.

For the Zetaproteobacteria, 28 OTUs have been defined. [4] Of interest were the two globally distributed OTUs that dominated the phylogenetic tree, two OTUs that seemed to originate in the deep subsurface, [10] and several endemic OTUs, along with the relatively limited detection of the isolated Zetaproteobacteria representative.

Classification

Zetaproteobacteria OTUs can now be classified according to the naming scheme used in McAllister et al. (2011). [4] The program ZetaHunter uses closed reference binning to identify sequences closely related to the established OTUs in addition to identifying novel Zetaproteobacteria OTUs. ZetaHunter's feature list continues to grow, but includes: 1) stable OTU binning, 2) sample comparison, 3) database and mask management options, 4) multi-threaded processing, 5) chimera checking, 6) checks for non-database-related sequences, and 7) OTU network maps. The ZetaHunter software can be downloaded at: https://github.com/mooreryan/ZetaHunter

Phylogeny of Mariprofundaceae by GTDB 07-RS207 [33] [34] [35]
Mariprofundaceae

Ghiorsea bivora Mori et al. 2017

Mariprofundus

" M. erugo " Garrison et al. 2019

M. ferrooxydans Emerson et al. 2010

" M. micogutta " Makita et al. 2017

" M. aestuarium " Chiu et al. 2017

" M. ferrinatatus " Chiu et al. 2017

Habitats

Ecological Niche

All of the habitats where Zetaproteobacteria have been found have (at least) two things in common: 1) they all provide an interface of steep redox gradients of oxygen and iron. [66] & 2) they are marine or brackish. [47]

Reduced hydrothermal fluids, for instance, exiting from vents in the deep-sea carry with them high concentrations of ferrous iron and other reduced chemical species, creating a gradient upward through a microbial mat of high- to low-ferrous iron. Similarly, oxygen from the overlying seawater diffuses into the microbial mat resulting in a downward gradient of high to low oxygen. Zetaproteobacteria are thought to live at the interface, where there is enough oxygen for use as an electron acceptor without there being too much oxygen for the organism to compete with the increased rate of chemical oxidation, and where there is enough ferrous iron for growth. [20] [66]

Iron oxidation is not always energetically favorable. Reference [42] discusses favorable conditions for iron oxidation in habitats that otherwise may have been thought to be dominated by the more energy yielding metabolisms of hydrogen or sulfur oxidation.

Note: Iron is not the only reduced chemical species accociated with these redox gradient environments. It is likely that Zetaproteobacteria are not all iron oxidizers.

Metabolism

Iron oxidation pathways in both acidophilic and circumneutral freshwater iron oxidation habitats, such as acid mine drainage or groundwater iron seeps, respectively, are better understood than marine circumneutral iron oxidation.

In recent years, researchers have made progress in suggesting possibilities for how the Zetaproteobacteria oxidize iron, primarily through comparative genomics. With this technique, genomes from organisms with similar function, for example the freshwater Fe-oxidizing Betaproteobacteria and the marine Fe-oxidizing Zetaproteobacteria, are compared to find genes that may be required for this function. Identifying the iron oxidation pathway in the Zetaproteobacteria began with the publication of the first described cultured representative, M. ferrooxydans strain PV-1. In this genome, the gene neighborhood of a molybdopterin oxidoreductase protein was identified as a place to start looking at candidate iron oxidation pathway genes. [67] In a follow-up analysis of a metagenomic sample, Singer et al. (2013) concluded that this molybdopterin oxidoreductase gene cassette was likely involved in Fe oxidation. [68] Comparative analysis of several single cell genomes, however, suggested an alternative conserved gene cassette with several cytochrome c and cytochrome oxidase genes to be involved in Fe oxidation. [69] For further reading on Fe oxidation pathways see reference. [70]

The phylogenetic distance between the Zetaproteobacteria and the Fe-oxidizing freshwater Betaproteobacteria suggests that Fe oxidation and the produced biominerals are the result of convergent evolution. [24] Comparative genomics has been able to identify several genes that are shared between the two clades, however, suggesting that the trait of Fe oxidation could have been horizontally transferred, possibly virally mediated. [71] [72]

Fe mats associated with the Zetaproteobacteria, in addition to oxidizing Fe have been found to have the genetic potential for denitrification, arsenic detoxification, Calvin-Benson-Bassham (CBB) cycle, and reductive tricarboxylic acid (rTCA) cycles. Novel primers have been designed to detect these genes in environmental samples. [73]

It is difficult at this point to speculate on the metabolism of the entire class of Zetaproteobacteria (with at least 28 different OTUs/species) with the limited sample size.

Suggested reviews

See also

Related Research Articles

<i>Riftia pachyptila</i> Giant tube worm (species of annelid)

Riftia pachyptila, commonly known as the giant tube worm and less commonly known as the giant beardworm, is a marine invertebrate in the phylum Annelida related to tube worms commonly found in the intertidal and pelagic zones. R. pachyptila lives on the floor of the Pacific Ocean near hydrothermal vents. The vents provide a natural ambient temperature in their environment ranging from 2 to 30 °C, and this organism can tolerate extremely high hydrogen sulfide levels. These worms can reach a length of 3 m, and their tubular bodies have a diameter of 4 cm (1.6 in).

<span class="mw-page-title-main">Kamaʻehuakanaloa Seamount</span> Active submarine volcano off the southeast coast of the island of Hawaii

Kamaʻehuakanaloa Seamount is an active submarine volcano about 22 mi (35 km) off the southeast coast of the island of Hawaii. The top of the seamount is about 3,200 ft (975 m) below sea level. This seamount is on the flank of Mauna Loa, the largest shield volcano on Earth. Kamaʻehuakanaloa is the newest volcano in the Hawaiian–Emperor seamount chain, a string of volcanoes that stretches about 3,900 mi (6,200 km) northwest of Kamaʻehuakanaloa. Unlike most active volcanoes in the Pacific Ocean that make up the active plate margins on the Pacific Ring of Fire, Kamaʻehuakanaloa and the other volcanoes of the Hawaiian–Emperor seamount chain are hotspot volcanoes and formed well away from the nearest plate boundary. Volcanoes in the Hawaiian Islands arise from the Hawaii hotspot, and as the youngest volcano in the chain, Kamaʻehuakanaloa is the only Hawaiian volcano in the deep submarine preshield stage of development.

<span class="mw-page-title-main">Sulfate-reducing microorganism</span> Microorganisms that "breathe" sulfates

Sulfate-reducing microorganisms (SRM) or sulfate-reducing prokaryotes (SRP) are a group composed of sulfate-reducing bacteria (SRB) and sulfate-reducing archaea (SRA), both of which can perform anaerobic respiration utilizing sulfate (SO2−
4
) as terminal electron acceptor, reducing it to hydrogen sulfide (H2S). Therefore, these sulfidogenic microorganisms "breathe" sulfate rather than molecular oxygen (O2), which is the terminal electron acceptor reduced to water (H2O) in aerobic respiration.

Ferroglobus is a genus of the Archaeoglobaceae.

<span class="mw-page-title-main">Iron cycle</span>

The iron cycle (Fe) is the biogeochemical cycle of iron through the atmosphere, hydrosphere, biosphere and lithosphere. While Fe is highly abundant in the Earth's crust, it is less common in oxygenated surface waters. Iron is a key micronutrient in primary productivity, and a limiting nutrient in the Southern ocean, eastern equatorial Pacific, and the subarctic Pacific referred to as High-Nutrient, Low-Chlorophyll (HNLC) regions of the ocean.

<span class="mw-page-title-main">Hydrogen cycle</span> Hydrogen exchange between the living and non-living world

The hydrogen cycle consists of hydrogen exchanges between biotic (living) and abiotic (non-living) sources and sinks of hydrogen-containing compounds.

Sulfur-reducing bacteria are microorganisms able to reduce elemental sulfur (S0) to hydrogen sulfide (H2S). These microbes use inorganic sulfur compounds as electron acceptors to sustain several activities such as respiration, conserving energy and growth, in absence of oxygen. The final product of these processes, sulfide, has a considerable influence on the chemistry of the environment and, in addition, is used as electron donor for a large variety of microbial metabolisms. Several types of bacteria and many non-methanogenic archaea can reduce sulfur. Microbial sulfur reduction was already shown in early studies, which highlighted the first proof of S0 reduction in a vibrioid bacterium from mud, with sulfur as electron acceptor and H
2
as electron donor. The first pure cultured species of sulfur-reducing bacteria, Desulfuromonas acetoxidans, was discovered in 1976 and described by Pfennig Norbert and Biebel Hanno as an anaerobic sulfur-reducing and acetate-oxidizing bacterium, not able to reduce sulfate. Only few taxa are true sulfur-reducing bacteria, using sulfur reduction as the only or main catabolic reaction. Normally, they couple this reaction with the oxidation of acetate, succinate or other organic compounds. In general, sulfate-reducing bacteria are able to use both sulfate and elemental sulfur as electron acceptors. Thanks to its abundancy and thermodynamic stability, sulfate is the most studied electron acceptor for anaerobic respiration that involves sulfur compounds. Elemental sulfur, however, is very abundant and important, especially in deep-sea hydrothermal vents, hot springs and other extreme environments, making its isolation more difficult. Some bacteria – such as Proteus, Campylobacter, Pseudomonas and Salmonella – have the ability to reduce sulfur, but can also use oxygen and other terminal electron acceptors.

<span class="mw-page-title-main">Iron-oxidizing bacteria</span> Bacteria deriving energy from dissolved iron

Iron-oxidizing bacteria are chemotrophic bacteria that derive energy by oxidizing dissolved iron. They are known to grow and proliferate in waters containing iron concentrations as low as 0.1 mg/L. However, at least 0.3 ppm of dissolved oxygen is needed to carry out the oxidation.

Lithotrophs are a diverse group of organisms using an inorganic substrate to obtain reducing equivalents for use in biosynthesis or energy conservation via aerobic or anaerobic respiration. While lithotrophs in the broader sense include photolithotrophs like plants, chemolithotrophs are exclusively microorganisms; no known macrofauna possesses the ability to use inorganic compounds as electron sources. Macrofauna and lithotrophs can form symbiotic relationships, in which case the lithotrophs are called "prokaryotic symbionts". An example of this is chemolithotrophic bacteria in giant tube worms or plastids, which are organelles within plant cells that may have evolved from photolithotrophic cyanobacteria-like organisms. Chemolithotrophs belong to the domains Bacteria and Archaea. The term "lithotroph" was created from the Greek terms 'lithos' (rock) and 'troph' (consumer), meaning "eaters of rock". Many but not all lithoautotrophs are extremophiles.

<i>Beggiatoa</i> Genus of bacteria

Beggiatoa is a genus of Gammaproteobacteria belonging to the order Thiotrichales, in the Pseudomonadota phylum. This genus was one of the first bacteria discovered by Ukrainian botanist Sergei Winogradsky. During his research in Anton de Bary's laboratory of botany in 1887, he found that  Beggiatoa  oxidized  hydrogen sulfide  (H2S) as an energy source, forming intracellular  sulfur  droplets, with oxygen as the terminal electron acceptor and CO2 used as a carbon source. Winogradsky named it in honor of the Italian doctor and botanist  Francesco Secondo Beggiato (1806 - 1883), from Venice. Winogradsky referred to this form of metabolism as "inorgoxidation" (oxidation of inorganic compounds), today called chemolithotrophy. These organisms live in sulfur-rich environments such as soil, both marine and freshwater, in the deep sea hydrothermal vents and in polluted marine environments. The finding represented the first discovery of lithotrophy. Two species of Beggiatoa have been formally described: the type species Beggiatoa alba and Beggiatoa leptomitoformis, the latter of which was only published in 2017. This colorless and filamentous bacterium, sometimes in association with other sulfur bacteria (for example the genus Thiothrix), can be arranged in biofilm visible to the naked eye formed by a very long white filamentous mat, the white color is due to the stored sulfur. Species of Beggiatoa have cells up to 200 µm in diameter and they are one of the largest prokaryotes on Earth.

<span class="mw-page-title-main">Gammaproteobacteria</span> Class of bacteria

Gammaproteobacteria is a class of bacteria in the phylum Pseudomonadota. It contains about 250 genera, which makes it the most genus-rich taxon of the Prokaryotes. Several medically, ecologically, and scientifically important groups of bacteria belong to this class. It is composed by all Gram-negative microbes and is the most phylogenetically and physiologically diverse class of Proteobacteria.

Biomining is the technique of extracting metals from ores and other solid materials typically using prokaryotes, fungi or plants. These organisms secrete different organic compounds that chelate metals from the environment and bring it back to the cell where they are typically used to coordinate electrons. It was discovered in the mid 1900s that microorganisms use metals in the cell. Some microbes can use stable metals such as iron, copper, zinc, and gold as well as unstable atoms such as uranium and thorium. Large chemostats of microbes can be grown to leach metals from their media. These vats of culture can then be transformed into many marketable metal compounds. Biomining is an environmentally friendly technique compared to typical mining. Mining releases many pollutants while the only chemicals released from biomining is any metabolites or gasses that the bacteria secrete. The same concept can be used for bioremediation models. Bacteria can be inoculated into environments contaminated with metals, oils, or other toxic compounds. The bacteria can clean the environment by absorbing these toxic compounds to create energy in the cell. Bacteria can mine for metals, clean oil spills, purify gold, and use radioactive elements for energy.

Hydrogen-oxidizing bacteria are a group of facultative autotrophs that can use hydrogen as an electron donor. They can be divided into aerobes and anaerobes. The former use hydrogen as an electron donor and oxygen as an acceptor while the latter use sulphate or nitrogen dioxide as electron acceptors. Species of both types have been isolated from a variety of environments, including fresh waters, sediments, soils, activated sludge, hot springs, hydrothermal vents and percolating water.

<i>Mariprofundus ferrooxydans</i> Species of bacterium

Mariprofundus ferrooxydans is a neutrophilic, chemolithotrophic, Gram-negative bacterium which can grow by oxidising ferrous to ferric iron. It is one of the few members of the class Zetaproteobacteria in the phylum Pseudomonadota. It is typically found in iron-rich deep sea environments, particularly at hydrothermal vents. M. ferrooxydans characteristically produces stalks of solid iron oxyhydroxides that form into iron mats. Genes that have been proposed to catalyze Fe(II) oxidation in M. ferrooxydans are similar to those involved in known metal redox pathways, and thus it serves as a good candidate for a model iron oxidizing organism.

Sulfurimonas is a bacterial genus within the class of Campylobacterota, known for reducing nitrate, oxidizing both sulfur and hydrogen, and containing Group IV hydrogenases. This genus consists of four species: Sulfurimonas autorophica, Sulfurimonas denitrificans, Sulfurimonas gotlandica, and Sulfurimonas paralvinellae. The genus' name is derived from "sulfur" in Latin and "monas" from Greek, together meaning a “sulfur-oxidizing rod”. The size of the bacteria varies between about 1.5-2.5 μm in length and 0.5-1.0 μm in width. Members of the genus Sulfurimonas are found in a variety of different environments which include deep sea-vents, marine sediments, and terrestrial habitats. Their ability to survive in extreme conditions is attributed to multiple copies of one enzyme. Phylogenetic analysis suggests that members of the genus Sulfurimonas have limited dispersal ability and its speciation was affected by geographical isolation rather than hydrothermal composition. Deep ocean currents affect the dispersal of Sulfurimonas spp., influencing its speciation. As shown in the MLSA report of deep-sea hydrothermal vents Campylobacterota, Sulfurimonas has a higher dispersal capability compared with deep sea hydrothermal vent thermophiles, indicating allopatric speciation.

Katrina Jane Edwards was a pioneering geomicrobiologist known for her studies of organisms living below the ocean floor, specifically exploring the interactions between the microbes and their geological surroundings, and how global processes were influenced by these interactions. She spearheaded the Center for Dark Energy Biosphere Investigation (C-DEBI) project at the University of Southern California, which is ongoing. Edwards also helped organize the deep biosphere research community by heading the Fe-Oxidizing Microbial Observatory Project on Loihi Seamount, and serving on several program steering committees involving ocean drilling. Edwards taught at the Woods Hole Oceanographic Institution (WHOI) and later became a professor at the University of Southern California.[1][2]

<span class="mw-page-title-main">Microbial oxidation of sulfur</span>

Microbial oxidation of sulfur is the oxidation of sulfur by microorganisms to build their structural components. The oxidation of inorganic compounds is the strategy primarily used by chemolithotrophic microorganisms to obtain energy to survive, grow and reproduce. Some inorganic forms of reduced sulfur, mainly sulfide (H2S/HS) and elemental sulfur (S0), can be oxidized by chemolithotrophic sulfur-oxidizing prokaryotes, usually coupled to the reduction of oxygen (O2) or nitrate (NO3). Anaerobic sulfur oxidizers include photolithoautotrophs that obtain their energy from sunlight, hydrogen from sulfide, and carbon from carbon dioxide (CO2).

<span class="mw-page-title-main">Hydrothermal vent microbial communities</span> Undersea unicellular organisms

The hydrothermal vent microbial community includes all unicellular organisms that live and reproduce in a chemically distinct area around hydrothermal vents. These include organisms in the microbial mat, free floating cells, or bacteria in an endosymbiotic relationship with animals. Chemolithoautotrophic bacteria derive nutrients and energy from the geological activity at Hydrothermal vents to fix carbon into organic forms. Viruses are also a part of the hydrothermal vent microbial community and their influence on the microbial ecology in these ecosystems is a burgeoning field of research.

<span class="mw-page-title-main">Vailuluʻu</span> Volcanic seamount in the Samoa Islands

Vailuluʻu is a volcanic seamount discovered in 1975. It rises from the sea floor to a depth of 593 m (1,946 ft) and is located between Taʻu and Rose islands at the eastern end of the Samoa hotspot chain. The basaltic seamount is considered to mark the current location of the Samoa hotspot. The summit of Vailuluʻu contains a 2 km wide, 400 m deep oval-shaped caldera. Two principal rift zones extend east and west from the summit, parallel to the trend of the Samoan hotspot. A third less prominent rift extends southeast of the summit.

<span class="mw-page-title-main">Eifuku</span> Two volcanic seamounts in the Northern Marianas

Eifuku and NW Eifuku (北西永福) are two seamounts in the Pacific Ocean. The better known one is NW Eifuku, where an unusual hydrothermal vent called "Champagne" produced droplets of liquid CO
2
. Both seamounts are located in the Northern Marianas and are volcanoes, part of the Izu-Bonin-Mariana Arc. NW Eifuku rises to 1,535 metres (5,036 ft) depth below sea level and is a 9 kilometres (5.6 mi) wide volcanic cone.

References

  1. 1 2 3 Emerson, D.; Rentz, J. A.; Lilburn, T. G.; Davis, R. E.; Aldrich, H.; Chan, C.; Moyer, C. L. (2007). Reysenbach, Anna-Louise (ed.). "A Novel Lineage of Proteobacteria Involved in Formation of Marine Fe-Oxidizing Microbial Mat Communities". PLOS ONE. 2 (8): e667. Bibcode:2007PLoSO...2..667E. doi: 10.1371/journal.pone.0000667 . PMC   1930151 . PMID   17668050.
  2. mariprofundus entry in LPSN ; Euzéby, J.P. (1997). "List of Bacterial Names with Standing in Nomenclature: a folder available on the Internet". International Journal of Systematic and Evolutionary Microbiology. 47 (2): 590–2. doi: 10.1099/00207713-47-2-590 . PMID   9103655.
  3. 1 2 3 Emerson, D.; Moyer, C. L. (2002). "Neutrophilic Fe-oxidizing bacteria are abundant at the Loihi Seamount hydrothermal vents and play a major role in Fe oxide deposition". Applied and Environmental Microbiology. 68 (6): 3085–3093. Bibcode:2002ApEnM..68.3085E. doi:10.1128/AEM.68.6.3085-3093.2002. PMC   123976 . PMID   12039770.
  4. 1 2 3 4 5 McAllister, S. M.; Davis, R. E.; McBeth, J. M.; Tebo, B. M.; Emerson, D.; Moyer, C. L. (2011). "Biodiversity and Emerging Biogeography of the Neutrophilic Iron-Oxidizing Zetaproteobacteria". Applied and Environmental Microbiology. 77 (15): 5445–5457. Bibcode:2011ApEnM..77.5445M. doi:10.1128/AEM.00533-11. PMC   3147450 . PMID   21666021.
  5. 1 2 Schauer, R.; Røy, H.; Augustin, N.; Gennerich, H. H.; Peters, M.; Wenzhoefer, F.; Amann, R.; Meyerdierks, A. (2011). "Bacterial sulfur cycling shapes microbial communities in surface sediments of an ultramafic hydrothermal vent field" (PDF). Environmental Microbiology. 13 (10): 2633–2648. Bibcode:2011EnvMi..13.2633S. doi:10.1111/j.1462-2920.2011.02530.x. PMID   21895907.
  6. 1 2 Hodges, T. W.; Olson, J. B. (2008). "Molecular Comparison of Bacterial Communities within Iron-Containing Flocculent Mats Associated with Submarine Volcanoes along the Kermadec Arc". Applied and Environmental Microbiology. 75 (6): 1650–1657. doi:10.1128/AEM.01835-08. PMC   2655482 . PMID   19114513.
  7. 1 2 Davis, R. E.; Stakes, D. S.; Wheat, C. G.; Moyer, C. L. (2009). "Bacterial Variability within an Iron-Silica-Manganese-rich Hydrothermal Mound Located Off-axis at the Cleft Segment, Juan de Fuca Ridge". Geomicrobiology Journal. 26 (8): 570–580. Bibcode:2009GmbJ...26..570D. doi:10.1080/01490450902889080. S2CID   53621745.
  8. 1 2 Forget, N. L.; Murdock, S. A.; Juniper, S. K. (2010). "Bacterial diversity in Fe-rich hydrothermal sediments at two South Tonga Arc submarine volcanoes". Geobiology. 8 (5): 417–432. Bibcode:2010Gbio....8..417F. doi:10.1111/j.1472-4669.2010.00247.x. PMID   20533949. S2CID   807482.
  9. 1 2 Handley, K. M.; Boothman, C.; Mills, R. A.; Pancost, R. D.; Lloyd, J. R. (2010). "Functional diversity of bacteria in a ferruginous hydrothermal sediment". The ISME Journal. 4 (9): 1193–1205. Bibcode:2010ISMEJ...4.1193H. doi: 10.1038/ismej.2010.38 . PMID   20410934.
  10. 1 2 3 Kato, S.; Yanagawa, K.; Sunamura, M.; Takano, Y.; Ishibashi, J. I.; Kakegawa, T.; Utsumi, M.; Yamanaka, T.; Toki, T.; Noguchi, T.; Kobayashi, K.; Moroi, A.; Kimura, H.; Kawarabayasi, Y.; Marumo, K.; Urabe, T.; Yamagishi, A. (2009). "Abundance of Zetaproteobacteria within crustal fluids in back-arc hydrothermal fields of the Southern Mariana Trough". Environmental Microbiology. 11 (12): 3210–3222. doi: 10.1111/j.1462-2920.2009.02031.x . PMID   19691504.
  11. 1 2 3 4 McBeth, J. M.; Little, B. J.; Ray, R. I.; Farrar, K. M.; Emerson, D. (2010). "Neutrophilic Iron-Oxidizing "Zetaproteobacteria" and Mild Steel Corrosion in Nearshore Marine Environments". Applied and Environmental Microbiology. 77 (4): 1405–1412. doi:10.1128/AEM.02095-10. PMC   3067224 . PMID   21131509.
  12. McBeth, J. M.; Fleming, E. J.; Emerson, D. (2013). "The transition from freshwater to marine iron-oxidizing bacterial lineages along a salinity gradient on the Sheepscot River, Maine, USA". Environmental Microbiology Reports. 5 (3): 453–63. doi:10.1111/1758-2229.12033. PMID   23754725.
  13. 1 2 3 McAllister, S. M.; Barnett, J. M.; Heiss, J. W.; Findlay, A. J.; MacDonald, D. J.; Dow, C. L.; Luther, G. W.; Michael, H. A.; Chan, C. S. (2015). "Dynamic hydrologic and biogeochemical processes drive microbially enhanced iron and sulfur cycling within the intertidal mixing zone of a beach aquifer". Limnology and Oceanography. 60 (1): 329–345. Bibcode:2015LimOc..60..329M. doi: 10.1002/lno.10029 .
  14. 1 2 Dang, H.; Chen, R.; Wang, L.; Shao, S.; Dai, L.; Ye, Y.; Guo, L.; Huang, G.; Klotz, M. G. (2011). "Molecular characterization of putative biocorroding microbiota with a novel niche detection of Epsilon- and Zetaproteobacteria in Pacific Ocean coastal seawaters". Environmental Microbiology. 13 (11): 3059–3074. doi:10.1111/j.1462-2920.2011.02583.x. PMID   21951343.
  15. Lee, J. S.; McBeth, J. M.; Ray, R. I.; Little, B. J.; Emerson, D. (2013). "Iron cycling at corroding carbon steel surfaces". Biofouling. 29 (10): 1243–1252. Bibcode:2013Biofo..29.1243L. doi:10.1080/08927014.2013.836184. PMC   3827670 . PMID   24093730.
  16. Moyer, C. L.; Dobbs, F. C.; Karl, D. M. (1995). "Phylogenetic diversity of the bacterial community from a microbial mat at an active, hydrothermal vent system, Loihi Seamount, Hawaii". Applied and Environmental Microbiology. 61 (4): 1555–1562. Bibcode:1995ApEnM..61.1555M. doi:10.1128/AEM.61.4.1555-1562.1995. PMC   167411 . PMID   7538279.
  17. Emerson, D.; Merrill Floyd, M. (2005). "Enrichment and Isolation of Iron‐Oxidizing Bacteria at Neutral pH". Environmental Microbiology. Methods in Enzymology. Vol. 397. pp. 112–23. doi:10.1016/S0076-6879(05)97006-7. ISBN   9780121828028. PMID   16260287.
  18. Summers, Z. M.; Gralnick, J. A.; Bond, D. R. (2013). "Cultivation of an Obligate Fe(II)-Oxidizing Lithoautotrophic Bacterium Using Electrodes". mBio. 4 (1): e00420–e00412. doi:10.1128/mBio.00420-12. PMC   3560526 . PMID   23362318.
  19. Barco, R. A.; Edwards, K. J. (2014). "Interactions of proteins with biogenic iron oxyhydroxides and a new culturing technique to increase biomass yields of neutrophilic, iron-oxidizing bacteria". Frontiers in Microbiology. 5: 259. doi: 10.3389/fmicb.2014.00259 . PMC   4038746 . PMID   24910632.
  20. 1 2 3 Emerson, D.; Fleming, E. J.; McBeth, J. M. (2010). "Iron-Oxidizing Bacteria: An Environmental and Genomic Perspective". Annual Review of Microbiology. 64: 561–583. doi:10.1146/annurev.micro.112408.134208. PMID   20565252.
  21. 1 2 3 Chan, C. S.; Fakra, S. C.; Emerson, D.; Fleming, E. J.; Edwards, K. J. (2010). "Lithotrophic iron-oxidizing bacteria produce organic stalks to control mineral growth: Implications for biosignature formation". The ISME Journal. 5 (4): 717–727. doi:10.1038/ismej.2010.173. PMC   3105749 . PMID   21107443.
  22. 1 2 Comolli, L. R.; Luef, B.; Chan, C. S. (2011). "High-resolution 2D and 3D cryo-TEM reveals structural adaptations of two stalk-forming bacteria to an Fe-oxidizing lifestyle". Environmental Microbiology. 13 (11): 2915–2929. doi:10.1111/j.1462-2920.2011.02567.x. PMID   21895918.
  23. Saini, G.; Chan, C. S. (2013). "Near-neutral surface charge and hydrophilicity prevent mineral encrustation of Fe-oxidizing micro-organisms". Geobiology. 11 (2): 191–200. Bibcode:2013Gbio...11..191S. doi:10.1111/gbi.12021. PMID   23279435. S2CID   205139903.
  24. 1 2 3 Fleming, E. J.; Davis, R. E.; McAllister, S. M.; Chan, C. S.; Moyer, C. L.; Tebo, B. M.; Emerson, D. (2013). "Hidden in plain sight: Discovery of sheath-forming, iron-oxidizing Zetaproteobacteriaat Loihi Seamount, Hawaii, USA". FEMS Microbiology Ecology. 85 (1): 116–127. Bibcode:2013FEMME..85..116F. doi: 10.1111/1574-6941.12104 . PMID   23480633.
  25. Juniper, S. Kim; Yves Fouquet (1988). "Filamentous iron-silica deposits from modern and ancient hydrothermal sites". Canadian Mineralogist. 26: 859–869.
  26. Hofmann, B. A.; Farmer, J. D.; von Blanckenburg, F.; Fallick, A. E. (2008). "Subsurface Filamentous Fabrics: An Evaluation of Origins Based on Morphological and Geochemical Criteria, with Implications for Exopaleontology". Astrobiology. 8 (1): 87–117. Bibcode:2008AsBio...8...87H. doi:10.1089/ast.2007.0130. PMID   18241094.
  27. Planavsky, N.; Rouxel, O.; Bekker, A.; Shapiro, R.; Fralick, P.; Knudsen, A. (2009). "Iron-oxidizing microbial ecosystems thrived in late Paleoproterozoic redox-stratified oceans". Earth and Planetary Science Letters. 286 (1–2): 230–242. Bibcode:2009E&PSL.286..230P. doi:10.1016/j.epsl.2009.06.033.
  28. Little, C. T. S.; Glynn, S. E. J.; Mills, R. A. (2004). "Four-Hundred-and-Ninety-Million-Year Record of Bacteriogenic Iron Oxide Precipitation at Sea-Floor Hydrothermal Vents" (PDF). Geomicrobiology Journal. 21 (6): 415–429. Bibcode:2004GmbJ...21..415L. doi:10.1080/01490450490485845. S2CID   54894240.
  29. Sun, Z.; Li, J.; Huang, W.; Dong, H.; Little, C. T. S.; Li, J. (2015). "Generation of hydrothermal Fe-Si oxyhydroxide deposit on the Southwest Indian Ridge and its implication for the origin of ancient banded iron formations" (PDF). Journal of Geophysical Research: Biogeosciences. 120 (1): 187–203. Bibcode:2015JGRG..120..187S. doi:10.1002/2014JG002764. S2CID   62895694.
  30. Krepski, S. T.; Emerson, D.; Hredzak-Showalter, P. L.; Luther, G. W.; Chan, C. S. (2013). "Morphology of biogenic iron oxides records microbial physiology and environmental conditions: Toward interpreting iron microfossils". Geobiology. 11 (5): 457–71. Bibcode:2013Gbio...11..457K. doi:10.1111/gbi.12043. PMID   23790206. S2CID   350074.
  31. Chan, C. S.; Fakra, S. C.; Edwards, D. C.; Emerson, D.; Banfield, J. F. (2009). "Iron oxyhydroxide mineralization on microbial extracellular polysaccharides". Geochimica et Cosmochimica Acta. 73 (13): 3807–3818. Bibcode:2009GeCoA..73.3807C. doi:10.1016/j.gca.2009.02.036. S2CID   31706001.
  32. Bennett, S. A.; Toner, B. M.; Barco, R.; Edwards, K. J. (2014). "Carbon adsorption onto Fe oxyhydroxide stalks produced by a lithotrophic iron-oxidizing bacteria". Geobiology. 12 (2): 146–156. Bibcode:2014Gbio...12..146B. doi:10.1111/gbi.12074. PMID   24428517. S2CID   205140009.
  33. "GTDB release 07-RS207". Genome Taxonomy Database . Retrieved 20 June 2022.
  34. "ar53_r207.sp_label". Genome Taxonomy Database . Retrieved 20 June 2022.
  35. "Taxon History". Genome Taxonomy Database . Retrieved 20 June 2022.
  36. Rassa, A. C.; McAllister, S. M.; Safran, S. A.; Moyer, C. L. (2009). "Zeta-Proteobacteria Dominate the Colonization and Formation of Microbial Mats in Low-Temperature Hydrothermal Vents at Loihi Seamount, Hawaii". Geomicrobiology Journal. 26 (8): 623–638. Bibcode:2009GmbJ...26..623R. doi:10.1080/01490450903263350. S2CID   53617751.
  37. Emerson, D.; Moyer, C. (2010). "Microbiology of Seamounts: Common Patterns Observed in Community Structure". Oceanography. 23: 148–163. doi: 10.5670/oceanog.2010.67 .
  38. Sudek, L. A.; Templeton, A. S.; Tebo, B. M.; Staudigel, H. (2009). "Microbial Ecology of Fe (hydr)oxide Mats and Basaltic Rock from Vailulu'u Seamount, American Samoa". Geomicrobiology Journal. 26 (8): 581–596. Bibcode:2009GmbJ...26..581S. doi:10.1080/01490450903263400. S2CID   85954222.
  39. Edwards, K. J.; Glazer, B. T.; Rouxel, O. J.; Bach, W.; Emerson, D.; Davis, R. E.; Toner, B. M.; Chan, C. S.; Tebo, B. M.; Staudigel, H.; Moyer, C. L. (2011). "Ultra-diffuse hydrothermal venting supports Fe-oxidizing bacteria and massive umber deposition at 5000 m off Hawaii". The ISME Journal. 5 (11): 1748–1758. Bibcode:2011ISMEJ...5.1748E. doi:10.1038/ismej.2011.48. PMC   3197161 . PMID   21544100.
  40. Kato, S.; Kobayashi, C.; Kakegawa, T.; Yamagishi, A. (2009). "Microbial communities in iron-silica-rich microbial mats at deep-sea hydrothermal fields of the Southern Mariana Trough". Environmental Microbiology. 11 (8): 2094–2111. doi:10.1111/j.1462-2920.2009.01930.x. PMID   19397679.
  41. Davis, R. E.; Moyer, C. L. (2008). "Extreme spatial and temporal variability of hydrothermal microbial mat communities along the Mariana Island Arc and southern Mariana back-arc system". Journal of Geophysical Research. 113 (B8): B08S15. Bibcode:2008JGRB..113.8S15D. doi: 10.1029/2007JB005413 .
  42. 1 2 Kato, S.; Nakamura, K.; Toki, T.; Ishibashi, J. I.; Tsunogai, U.; Hirota, A.; Ohkuma, M.; Yamagishi, A. (2012). "Iron-Based Microbial Ecosystem on and Below the Seafloor: A Case Study of Hydrothermal Fields of the Southern Mariana Trough". Frontiers in Microbiology. 3: 89. doi: 10.3389/fmicb.2012.00089 . PMC   3304087 . PMID   22435065.
  43. Meyer-Dombard, D. A. R.; Amend, J. P.; Osburn, M. R. (2012). "Microbial diversity and potential for arsenic and iron biogeochemical cycling at an arsenic rich, shallow-sea hydrothermal vent (Tutum Bay, Papua New Guinea)". Chemical Geology. 348: 37–47. doi:10.1016/j.chemgeo.2012.02.024.
  44. Li, J.; Zhou, H.; Peng, X.; Wu, Z.; Chen, S.; Fang, J. (2012). "Microbial diversity and biomineralization in low-temperature hydrothermal iron-silica-rich precipitates of the Lau Basin hydrothermal field". FEMS Microbiology Ecology. 81 (1): 205–216. Bibcode:2012FEMME..81..205L. doi: 10.1111/j.1574-6941.2012.01367.x . PMID   22443540.
  45. Dekov, V. M.; Petersen, S.; Garbe-Schönberg, C. -D.; Kamenov, G. D.; Perner, M.; Kuzmann, E.; Schmidt, M. (2010). "Fe–Si-oxyhydroxide deposits at a slow-spreading centre with thickened oceanic crust: The Lilliput hydrothermal field (9°33′S, Mid-Atlantic Ridge)". Chemical Geology. 278 (3–4): 186–200. Bibcode:2010ChGeo.278..186D. doi:10.1016/j.chemgeo.2010.09.012.
  46. 1 2 3 MacDonald, D. J.; Findlay, A. J.; McAllister, S. M.; Barnett, J. M.; Hredzak-Showalter, P.; Krepski, S. T.; Cone, S. G.; Scott, J.; Bennett, S. K.; Chan, C. S.; Emerson, D.; Luther Iii, G. W. (2014). "Using in situ voltammetry as a tool to identify and characterize habitats of iron-oxidizing bacteria: From fresh water wetlands to hydrothermal vent sites". Environmental Science: Processes & Impacts. 16 (9): 2117–2126. doi:10.1039/c4em00073k. PMID   24924809.
  47. 1 2 Scott, J. J.; Breier, J. A.; Luther, G. W.; Emerson, D. (2015). "Microbial Iron Mats at the Mid-Atlantic Ridge and Evidence that Zetaproteobacteria May Be Restricted to Iron-Oxidizing Marine Systems". PLOS ONE. 10 (3): e0119284. Bibcode:2015PLoSO..1019284S. doi: 10.1371/journal.pone.0119284 . PMC   4356598 . PMID   25760332.
  48. Cao, H.; Wang, Y.; Lee, O. O.; Zeng, X.; Shao, Z.; Qian, P. -Y. (2014). "Microbial Sulfur Cycle in Two Hydrothermal Chimneys on the Southwest Indian Ridge". mBio. 5 (1): e00980–e00913. doi:10.1128/mBio.00980-13. PMC   3903282 . PMID   24473131.
  49. Sylvan, J. B.; Toner, B. M.; Edwards, K. J. (2012). "Life and Death of Deep-Sea Vents: Bacterial Diversity and Ecosystem Succession on Inactive Hydrothermal Sulfides". mBio. 3 (1): e00279–e00211. doi:10.1128/mBio.00279-11. PMC   3262234 . PMID   22275502.
  50. Dhillon, A.; Teske, A.; Dillon, J.; Stahl, D. A.; Sogin, M. L. (2003). "Molecular characterization of sulfate-reducing bacteria in the Guaymas Basin". Applied and Environmental Microbiology. 69 (5): 2765–2772. Bibcode:2003ApEnM..69.2765D. doi:10.1128/AEM.69.5.2765-2772.2003. PMC   154542 . PMID   12732547.
  51. Kato, S.; Ikehata, K.; Shibuya, T.; Urabe, T.; Ohkuma, M.; Yamagishi, A. (2015). "Potential for biogeochemical cycling of sulfur, iron and carbon within massive sulfide deposits below the seafloor". Environmental Microbiology. 17 (5): 1817–35. Bibcode:2015EnvMi..17.1817K. doi:10.1111/1462-2920.12648. PMID   25330135.
  52. Jacobson Meyers, M. E.; Sylvan, J. B.; Edwards, K. J. (2014). "Extracellular Enzyme Activity and Microbial Diversity Measured on Seafloor Exposed Basalts from Loihi Seamount Indicate the Importance of Basalts to Global Biogeochemical Cycling". Applied and Environmental Microbiology. 80 (16): 4854–64. Bibcode:2014ApEnM..80.4854J. doi:10.1128/AEM.01038-14. PMC   4135773 . PMID   24907315.
  53. Rubin-Blum, M.; Antler, G.; Tsadok, R.; Shemesh, E.; Austin, J. A.; Coleman, D. F.; Goodman-Tchernov, B. N.; Ben-Avraham, Z.; Tchernov, D. (2014). "First Evidence for the Presence of Iron Oxidizing Zetaproteobacteria at the Levantine Continental Margins". PLOS ONE. 9 (3): e91456. Bibcode:2014PLoSO...991456R. doi: 10.1371/journal.pone.0091456 . PMC   3948872 . PMID   24614177.
  54. Bowman, J. P.; McCuaig, R. D. (2003). "Biodiversity, community structural shifts, and biogeography of prokaryotes within Antarctic continental shelf sediment". Applied and Environmental Microbiology. 69 (5): 2463–2483. Bibcode:2003ApEnM..69.2463B. doi:10.1128/AEM.69.5.2463-2483.2003. PMC   154503 . PMID   12732511.
  55. Eder, W.; Jahnke, L. L.; Schmidt, M.; Huber, R. (2001). "Microbial Diversity of the Brine-Seawater Interface of the Kebrit Deep, Red Sea, Studied via 16S rRNA Gene Sequences and Cultivation Methods". Applied and Environmental Microbiology. 67 (7): 3077–3085. Bibcode:2001ApEnM..67.3077E. doi:10.1128/AEM.67.7.3077-3085.2001. PMC   92984 . PMID   11425725.
  56. Moreau, J. W.; Zierenberg, R. A.; Banfield, J. F. (2010). "Diversity of Dissimilatory Sulfite Reductase Genes (dsrAB) in a Salt Marsh Impacted by Long-Term Acid Mine Drainage". Applied and Environmental Microbiology. 76 (14): 4819–4828. Bibcode:2010ApEnM..76.4819M. doi:10.1128/AEM.03006-09. PMC   2901737 . PMID   20472728.
  57. Stauffert, M.; Cravo-Laureau, C.; Jézéquel, R.; Barantal, S.; Cuny, P.; Gilbert, F.; Cagnon, C.; Militon, C. C.; Amouroux, D.; Mahdaoui, F.; Bouyssiere, B.; Stora, G.; Merlin, F. O. X.; Duran, R. (2013). "Impact of Oil on Bacterial Community Structure in Bioturbated Sediments". PLOS ONE. 8 (6): e65347. Bibcode:2013PLoSO...865347S. doi: 10.1371/journal.pone.0065347 . PMC   3677869 . PMID   23762350.
  58. Pischedda, L.; Militon, C. C.; Gilbert, F.; Cuny, P. (2011). "Characterization of specificity of bacterial community structure within the burrow environment of the marine polychaete Hediste (Nereis) diversicolor". Research in Microbiology. 162 (10): 1033–42. doi: 10.1016/j.resmic.2011.07.008 . PMID   21946148. S2CID   11103232.
  59. Asano, R.; Nakai, Y.; Kawada, W.; Shimura, Y.; Inamoto, T.; Fukushima, J. (2013). "Seawater Inundation from the 2011 Tohoku Tsunami Continues to Strongly Affect Soil Bacterial Communities 1 Year Later". Microbial Ecology. 66 (3): 639–46. Bibcode:2013MicEc..66..639A. doi:10.1007/s00248-013-0261-9. PMID   23846833. S2CID   16593799.
  60. Thompson, C.; Beys-Da-Silva, W.; Santi, L. L.; Berger, M.; Vainstein, M.; Guima Rães, J.; Vasconcelos, A. T. (2013). "A potential source for cellulolytic enzyme discovery and environmental aspects revealed through metagenomics of Brazilian mangroves". AMB Express. 3 (1): 65. doi: 10.1186/2191-0855-3-65 . PMC   3922913 . PMID   24160319.
  61. Emerson, J. B.; Thomas, B. C.; Alvarez, W.; Banfield, J. F. (2015). "Metagenomic analysis of a high carbon dioxide subsurface microbial community populated by chemolithoautotrophs and bacteria and archaea from candidate phyla". Environmental Microbiology. 18 (6): 1686–1703. doi:10.1111/1462-2920.12817. OSTI   1571927. PMID   25727367. S2CID   4677869.
  62. Colman, D. R.; Garcia, J. R.; Crossey, L. J.; Karlstrom, K.; Jackson-Weaver, O.; Takacs-Vesbach, C. (2014). "An analysis of geothermal and carbonic springs in the western United States sustained by deep fluid inputs". Geobiology. 12 (1): 83–98. Bibcode:2014Gbio...12...83C. doi:10.1111/gbi.12070. PMID   24286205. S2CID   22057647.
  63. Ionescu, D.; Heim, C.; Polerecky, L.; Ramette, A.; Haeusler, S.; Bizic-Ionescu, M.; Thiel, V.; De Beer, D. (2015). "Diversity of Iron Oxidizing and Reducing Bacteria in Flow Reactors in the Äspö Hard Rock Laboratory". Geomicrobiology Journal. 32 (3–4): 207–220. Bibcode:2015GmbJ...32..207I. doi:10.1080/01490451.2014.884196. S2CID   96387274.
  64. Zbinden, M.; Cambon-Bonavita, M. A. (2003). "Occurrence of Deferribacterales and Entomoplasmatales in the deep-sea Alvinocarid shrimp Rimicaris exoculata gut". FEMS Microbiology Ecology. 46 (1): 23–30. Bibcode:2003FEMME..46...23Z. doi: 10.1016/S0168-6496(03)00176-4 . PMID   19719579.
  65. Jan, C.; Petersen, J. M.; Werner, J.; Teeling, H.; Huang, S.; Glöckner, F. O.; Golyshina, O. V.; Dubilier, N.; Golyshin, P. N.; Jebbar, M.; Cambon-Bonavita, M. A. (2014). "The gill chamber epibiosis of deep-sea shrimp Rimicarisexoculata: An in-depth metagenomic investigation and discovery of Zetaproteobacteria". Environmental Microbiology. 16 (9): 2723–38. doi:10.1111/1462-2920.12406. PMID   24447589.
  66. 1 2 Glazer, B. T.; Rouxel, O. J. (2009). "Redox Speciation and Distribution within Diverse Iron-dominated Microbial Habitats at Loihi Seamount". Geomicrobiology Journal. 26 (8): 606–622. Bibcode:2009GmbJ...26..606G. doi:10.1080/01490450903263392. S2CID   50934359.
  67. Singer, E.; Emerson, D.; Webb, E. A.; Barco, R. A.; Kuenen, J. G.; Nelson, W. C.; Chan, C. S.; Comolli, L. R.; Ferriera, S.; Johnson, J.; Heidelberg, J. F.; Edwards, K. J. (2011). Khodursky, Arkady B (ed.). "Mariprofundus ferrooxydans PV-1 the First Genome of a Marine Fe(II) Oxidizing Zetaproteobacterium". PLOS ONE. 6 (9): e25386. Bibcode:2011PLoSO...625386S. doi: 10.1371/journal.pone.0025386 . PMC   3179512 . PMID   21966516.
  68. Singer, E.; Heidelberg, J. F.; Dhillon, A.; Edwards, K. J. (2013). "Metagenomic insights into the dominant Fe(II) oxidizing Zetaproteobacteria from an iron mat at Lō´ihi, Hawai´l". Frontiers in Microbiology. 4: 52. doi: 10.3389/fmicb.2013.00052 . PMC   3603346 . PMID   23518919.
  69. Field, E. K.; Sczyrba, A.; Lyman, A. E.; Harris, C. C.; Woyke, T.; Stepanauskas, R.; Emerson, D. (2014). "Genomic insights into the uncultivated marine Zetaproteobacteria at Loihi Seamount". The ISME Journal. 9 (4): 857–70. doi:10.1038/ismej.2014.183. PMC   4817698 . PMID   25303714.
  70. 1 2 Ilbert, M.; Bonnefoy, V. (2013). "Insight into the evolution of the iron oxidation pathways". Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1827 (2): 161–175. doi: 10.1016/j.bbabio.2012.10.001 . PMID   23044392.
  71. Emerson, D.; Field, E. K.; Chertkov, O.; Davenport, K. W.; Goodwin, L.; Munk, C.; Nolan, M.; Woyke, T. (2013). "Comparative genomics of freshwater Fe-oxidizing bacteria: Implications for physiology, ecology, and systematics". Frontiers in Microbiology. 4: 254. doi: 10.3389/fmicb.2013.00254 . PMC   3770913 . PMID   24062729.
  72. Altermann, E. (2014). "Invited commentary: Lubricating the rusty wheel, new insights into iron oxidizing bacteria through comparative genomics". Frontiers in Microbiology. 5: 386. doi: 10.3389/fmicb.2014.00386 . PMC   4115626 . PMID   25126088.
  73. Jesser, K. J.; Fullerton, H.; Hager, K. W.; Moyer, C. L. (2015). "Quantitative PCR Analysis of Functional Genes in Iron-Rich Microbial Mats at an Active Hydrothermal Vent System (Lō'ihi Seamount, Hawai'i)". Applied and Environmental Microbiology. 81 (9): 2976–2984. Bibcode:2015ApEnM..81.2976J. doi:10.1128/AEM.03608-14. PMC   4393452 . PMID   25681182.
  74. Hedrich, S.; Schlomann, M.; Johnson, D. B. (2011). "The iron-oxidizing proteobacteria". Microbiology. 157 (6): 1551–1564. doi: 10.1099/mic.0.045344-0 . PMID   21511765.
  75. Kato, Shingo (2015). "Ecophysiology of neutrophilic iron-oxidizing microorganisms and its significance in global biogeochemical cycling". Chikyukagaku (Geochemistry). 49: 1–17. doi:10.14934/chikyukagaku.49.1.
  76. Ishibashi, Jun-Ichiro; Okino, Kyoko; Sunamura, Michinari, eds. (2015). Subseafloor Biosphere Linked to Hydrothermal Systems. doi:10.1007/978-4-431-54865-2. ISBN   978-4-431-54864-5. S2CID   133245369.
  77. Melton, E. D.; Swanner, E. D.; Behrens, S.; Schmidt, C.; Kappler, A. (2014). "The interplay of microbially mediated and abiotic reactions in the biogeochemical Fe cycle". Nature Reviews Microbiology. 12 (12): 797–808. doi:10.1038/nrmicro3347. PMID   25329406. S2CID   24058676.