Microbial oxidation of sulfur

Last updated
Reactions of oxidation of sulfide to sulfate and elemental sulfur (incorrectly balanced). The electrons (e ) liberated from these oxidation reactions, which release chemical energy, are then used to fix carbon into organic molecules. The elements that become oxidized are shown in pink, those that become reduced in blue, and the electrons in purple. Sulfide oxidation reactions.jpg
Reactions of oxidation of sulfide to sulfate and elemental sulfur (incorrectly balanced). The electrons (e ) liberated from these oxidation reactions, which release chemical energy, are then used to fix carbon into organic molecules. The elements that become oxidized are shown in pink, those that become reduced in blue, and the electrons in purple.

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). [1] [2] Anaerobic sulfur oxidizers include photolithoautotrophs that obtain their energy from sunlight, hydrogen from sulfide, and carbon from carbon dioxide (CO2).

Contents

Most of the sulfur oxidizers are autotrophs that can use reduced sulfur species as electron donors for CO2 fixation. The microbial oxidation of sulfur is an important link in the biogeochemical cycling of sulfur in environments hosting both abundant reduced sulfur species and low concentrations of oxygen, such as marine sediments, oxygen minimum zones (OMZs) and hydrothermal systems. [3]

Ecology

The oxidation of hydrogen sulfide has been considered one of the most important processes in the environment, given that the oceans have had very low oxygen and high sulfidic conditions over most of the Earth's history. The modern analog ecosystems are deep marine basins, for instance in the Black Sea, near the Cariaco trench and the Santa Barbara basin. Other zones of the ocean that experience periodic anoxic and sulfidic conditions are the upwelling zones off the coasts of Chile and Namibia, and hydrothermal vents, which are a key source of H2S to the ocean. [4] Sulfur oxidizing microorganisms (SOM) are thus restricted to upper sediment layers in these environments, where oxygen and nitrate are available. The SOM may play an important yet unconsidered role in carbon sequestration, [5] since some models [6] and experiments with Gammaproteobacteria [7] [8] have suggested that sulfur-dependent carbon fixation in marine sediments could be responsible for almost half of total dark carbon fixation in the oceans. Besides, they may have been critical for the evolution of eukaryotic organisms, given that sulfur metabolism could have driven the formation of the symbiotic associations that sustained them [9] (see below).

Although the biological oxidation of reduced sulfur compounds competes with abiotic chemical reactions (e.g. the iron-mediated oxidation of sulfide to iron sulfide (FeS) or pyrite (FeS2)), [10] thermodynamic and kinetic considerations suggest that biological oxidation far exceeds the chemical oxidation of sulfide in most environments. [4] Experimental data from the anaerobic phototroph Chlorobaculum tepidum indicate that microorganisms enhance sulfide oxidation by three or more orders of magnitude. [4] However, the general contribution of microorganisms to total sulfur oxidation in marine sediments is still unknown. The SOM of Alphaproteobacteria, Gammaproteobacteria and Campylobacterota account for average cell abundances of 108 cells/m3 in organic-rich marine sediments. [11] Considering that these organisms have a very narrow range of habitats, as explained below, a major fraction of sulfur oxidation in many marine sediments may be accounted for by these groups. [12]

Given that the maximal concentrations of oxygen, nitrate and sulfide are usually separated in depth profiles, many SOM cannot directly access their hydrogen or electron sources (reduced sulfur species) and energy sources (O2 or nitrate) at the same time. This limitation has led SOM to develop different morphological adaptations. [12] The large sulfur bacteria (LSB) of the family Beggiatoaceae (Gammaproteobacteria) have been used as model organisms for benthic sulfur oxidation. They are known as 'gradient organisms' that are indicative of hypoxic (low oxygen) and sulfidic (rich in reduced sulfur species) conditions. They internally store large amounts of nitrate and elemental sulfur to overcome the spatial gap between oxygen and sulfide. Some of the Beggiatoaceae are filamentous and can thus glide between oxic/suboxic and sulfidic environments, while the non-motile ones rely on nutrient suspensions, fluxes, or attach themselves to bigger particles. [12] Some marine non-motile LSB are the only known free-living bacteria that have two carbon fixation pathways: the Calvin-Benson cycle (used by plants and other photosynthetic organisms) and the reverse tricarboxylic acid cycle. [13]

Another evolutionary strategy of SOM is to partner up with motile eukaryotic organisms. The symbiotic SOM provides carbon and, in some cases, bioavailable nitrogen to the host, and gets enhanced access to resources and shelter in return. This lifestyle has evolved independently in sediment-dwelling ciliates, oligochaetes, nematodes, flatworms and bivalves. [14] Recently, a new mechanism for sulfur oxidation was discovered in filamentous bacteria. It is called electrogenic sulfur oxidation (e-SOx), and involves the formation of multicellular bridges that connect the oxidation of sulfide in anoxic sediment layers with the reduction of oxygen or nitrate in oxic surface sediments, generating electric currents over centimeter distances. The so-called cable bacteria are widespread in shallow marine sediments, [15] and are believed to conduct electrons through structures inside a common periplasm of the multicellular filament, [16] a process that may influence the cycling of elements at aquatic sediment surfaces, for instance, by altering iron speciation. [17] The LSB and cable bacteria seem to be restricted to undisturbed sediment with stable hydrodynamic conditions, [18] while symbiotic SOM and their hosts have been mainly found in permeable coastal sediments. [12]

Microbial diversity

The oxidation of reduced sulfur compounds is performed exclusively by Bacteria and Archaea. All the Archaea involved in this process are aerobic and belong to the Order Sulfolobales, [19] [20] characterized by acidophiles (extremophiles that require low pHs to grow) and thermophiles (extremophiles that require high temperatures to grow). The most studied have been the genera Sulfolobus, an aerobic Archaea, and Acidianus, a facultative anaerobe (i.e. an organism that can obtain energy either by aerobic or anaerobic respiration).

Sulfur oxidizing bacteria (SOB) are aerobic, anaerobic or facultative, and most of them are obligate or facultative autotrophs that can use either carbon dioxide or organic compounds as a source of carbon (mixotrophs). [3] The most abundant and studied SOB are in the family Thiobacilliaceae in terrestrial environments, and in the family Beggiatoaceae in aquatic environments. [3] Aerobic sulfur oxidizing bacteria are mainly mesophilic, growing at moderate ranges of temperature and pH, although some are thermophilic and/or acidophilic. Outside these families, other SOB described belong to the genera Acidithiobacillus , [21] Aquaspirillum , [22] Aquifex , [23] Bacillus , [24] Methylobacterium , [25] Paracoccus, Pseudomonas [22] Starkeya , [26] Thermithiobacillus , [21] and Xanthobacter . [22] On the other hand, the cable bacteria belong to the family Desulfobulbaceae of the Deltaproteobacteria and are currently represented by two candidate Genera, "Candidatus Electronema" and "Candidatus Electrothrix" [27] .

Anaerobic SOB (AnSOB) are mainly neutrophilic/mesophilic photosynthetic autotrophs, obtaining energy from sunlight but using reduced sulfur compounds instead of water as hydrogen or electron donors for photosynthesis. AnSOB include some purple sulfur bacteria (Chromatiaceae) [28] such as Allochromatium, [29] and green sulfur bacteria (Chlorobiaceae), as well as the purple non-sulfur bacteria (Rhodospirillaceae) [30] and some Cyanobacteria. [3] The AnSOB Cyanobacteria are only able to oxidize sulfide to elemental sulfur and have been identified as Oscillatoria, Lyngbya, Aphanotece, Microcoleus, and Phormidium. [31] [32] Some AnSOB, such as the facultative anaerobes Thiobacillus spp., and Thermothrix sp., are chemolithoautotrophs, meaning that they obtain energy from the oxidation of reduced sulfur species, which is then used to fix CO2. Others, such as some filamentous gliding green bacteria (Chloroflexaceae), are mixotrophs. From all of the SOB, the only group that directly oxidize sulfide to sulfate in abundance of oxygen without accumulating elemental sulfur are the Thiobacilli. The other groups accumulate elemental sulfur, which they may oxidize to sulfate when sulfide is limited or depleted. [3]

Biochemistry

Enzymatic pathways used by sulfide-oxidizing microorganisms. Left: SQR pathway. Right: Sox pathway. HS : sulfide; S : elemental sulfur; SO3 : sulfite; APS: adenosine-5'-phosphosulfate; SO4 : sulfate. Redrawn (adapted) with permission from Poser, A., Vogt, C., Knoller, K., Ahlheim, J., Weiss, H., Kleinsteuber, S., & Richnow, H. H. (2014). Stable sulfur and oxygen isotope fractionation of anoxic sulfide oxidation by two different enzymatic pathways. Environmental Science & Technology, 48(16), 9094-9102. Copyright 2008 American Chemical Society. Sulfide oxidation pathways.jpg
Enzymatic pathways used by sulfide-oxidizing microorganisms. Left: SQR pathway. Right: Sox pathway. HS : sulfide; S : elemental sulfur; SO3 : sulfite; APS: adenosine-5'-phosphosulfate; SO4 : sulfate. Redrawn (adapted) with permission from Poser, A., Vogt, C., Knöller, K., Ahlheim, J., Weiss, H., Kleinsteuber, S., & Richnow, H. H. (2014). Stable sulfur and oxygen isotope fractionation of anoxic sulfide oxidation by two different enzymatic pathways. Environmental Science & Technology, 48(16), 9094–9102. Copyright 2008 American Chemical Society.

There are two described pathways for the microbial oxidation of sulfide:

Similarly, two pathways for the oxidation of sulfite (SO32-) have been identified:

On the other hand, at least three pathways exist for the oxidation of thiosulfate (S2O32-) :

In any of these pathways, oxygen is the preferred electron acceptor, but in oxygen-limited environments nitrate, oxidized forms of iron and even organic matter are used instead. [44]

Cyanobacteria normally perform oxygenic photosynthesis using water as electron donor. However, in the presence of sulfide, oxygenic photosynthesis is inhibited, and some cyanobacteria can perform anoxygenic photosynthesis by oxidation of sulfide to thiosulfate − using Photosystem I with sulfite− as a possible intermediate sulfur compound. [45] [46]

Oxidation of sulfide

Sulfide oxidation can proceed under aerobic or anaerobic conditions. Aerobic sulfide-oxidizing bacteria usually oxidize sulfide to sulfate and are obligate or facultative chemolithoautotrophs. The latter can grow as heterotrophs, obtaining carbon from organic sources, or as autotrophs, using sulfide as the electron donor (energy source) for CO2 fixation. [3] The oxidation of sulfide can proceed aerobically by two different mechanisms: substrate-level phosphorylation, which is dependent on adenosine monophosphate (AMP), and oxidative phosphorylation independent of AMP, [47] which has been detected in several Thiobacilli (T. denitrificans, T. thioparus, T. novellus and T. neapolitanus), as well as in Acidithiobacillus ferrooxidans. [48] The archaeon Acidianus ambivalens appears to possess both an ADP-dependent and an ADP independent pathway for the oxidation of sulfide. [49] Similarly, both mechanisms operate in the chemoautotroph Thiobacillus denitrificans, [50] which can oxidize sulfide to sulfate anaerobically using nitrate as terminal electron acceptor [51] which is reduced to dinitrogen (N2). [52] Two other anaerobic strains that can perform a similar process were identified as similar to Thiomicrospira denitrificans and Arcobacter. [53]

Among the heterotrophic SOB are included species of Beggiatoa that can grow mixotrophically, using sulfide to obtain energy (autotrophic metabolism) or to eliminate metabolically formed hydrogen peroxide in the absence of catalase (heterotrophic metabolism). [54] Other organisms, such as the Bacteria Sphaerotilus natans [55] and the yeast Alternaria [56] are able to oxidize sulfide to elemental sulfur by means of the rDsr pathway. [57]

Oxidation of elemental sulfur

Some Bacteria and Archaea can aerobically oxidize elemental sulfur to sulfuric acid. [3] Acidithiobacillus ferrooxidans and Thiobacillus thioparus can oxidize sulfur to sulfite by means of an oxygenase enzyme, although it is thought that an oxidase could be used as well as an energy saving mechanism. [58] For the anaerobic oxidation of elemental sulfur, it is thought that the Sox pathway plays an important role, although this is not yet completely understood. [39] Thiobacillus denitrificans uses oxidized forms on nitrogen as energy source and terminal electron acceptor instead of oxygen. [59]

Oxidation of thiosulfate and tetrathionate

Most of the chemosynthetic autotrophic bacteria that can oxidize elemental sulfur to sulfate are also able to oxidize thiosulfate to sulfate as a source of reducing power for carbon dioxide assimilation. However, the mechanisms that they utilize may vary, since some of them, such as the photosynthetic purple bacteria, transiently accumulate extracellular elemental sulfur during the oxidation of tetrathionate before oxidizing it to sulfate, while the green sulfur bacteria do not. [3] A direct oxidation reaction (T. versutus [60] ), as well as others that involve sulfite (T. denitrificans) and tetrathionate (A. ferrooxidans, A. thiooxidans and Acidiphilum acidophilum [61] ) as intermediate compounds, have been proposed. Some mixotrophic bacteria only oxidize thiosulfate to tetrathionate. [3]

The mechanism of bacterial oxidation of tetrathionate is still unclear and may involve sulfur disproportionation, during which both sulfide and sulfate are produced from reduced sulfur species, and hydrolysis reactions. [3]

Isotope fractionations

The fractionation of sulfur and oxygen isotopes during microbial sulfide oxidation (MSO) has been studied to assess its potential as a proxy to differentiate it from the abiotic oxidation of sulfur. [62] The light isotopes of the elements that are most commonly found in organic molecules, such as 12C, 16O, 1H, 14N and 32S, form bonds that are broken slightly more easily than bonds between the corresponding heavy isotopes, 13C, 18O, 2H, 15N and 34S . Because there is a lower energetic cost associated with the use of light isotopes, enzymatic processes usually discriminate against the heavy isotopes, and, as a consequence, biological fractionations of isotopes are expected between the reactants and the products. A normal kinetic isotope effect is that in which the products are depleted in the heavy isotope relative to the reactants (low heavy isotope to light isotope ratio), and although this is not always the case, the study of isotope fractionations between enzymatic processes may allow tracing the source of the product.

Fractionation of oxygen isotopes

The formation of sulfate in aerobic conditions entails the incorporation of four oxygen atoms from water, and when coupled with dissimilatory nitrate reduction (DNR) -the preferential reduction pathway under anoxic conditions- can have a contribution of oxygen atoms from nitrate as well. The δ18O value of the newly formed sulfate thus depends on the δ18O value of the water, the isotopic fractionation associated with the incorporation of oxygen atoms from water to sulfate and a potential exchange of oxygen atoms between sulfur and nitrogen intermediates and water. [63] MSO has been found to produce small fractionations in 18O compared to water (~5‰). Given the very small fractionation of 18O that usually accompanies MSO, the relatively higher depletions in 18O of the sulfate produced by MSO coupled to DNR (-1.8 to -8.5 ‰) suggest a kinetic isotope effect in the incorporation of oxygen from water to sulfate and the role of nitrate as a potential alternative source of light oxygen. [63] The fractionations of oxygen produced by sulfur disproportionation from elemental sulfur have been found to be higher, with reported values from 8 to 18.4‰, which suggests a kinetic isotope effect in the pathways involved in oxidation of elemental sulfur to sulfate, although more studies are necessary to determine what are the specific steps and conditions that favor this fractionation. The table below summarizes the reported fractionations of oxygen isotopes from MSO in different organisms and conditions.

Starting compound (reactant)Intermediate or end compounds
(products)
OrganismAverage 18O fractionation (product/reactant)DetailsReference
SulfideSulfateA. ferrooxidans (chemolithotroph)4.1‰ (30 °C)AerobicTaylor et al. (1984) [64]
A. ferrooxidans (chemolithotroph)6.4‰
3.8‰

(no temperature provided)

Aerobic

Anaerobic

Thurston et al. (2010) [65]
Thiomicrospira sp. strain CVO (chemolithotroph)0‰

(no temperature provided)

Anaerobic, coupled to DNRHubert et al. (2009) [66]
T. denitrificans (chemolithotroph)
Sulfurimonas denitrificans

(chemolithotroph)

−6 to −1.8‰ (30 °C)


−8.5 to −2.1‰ (21 °C)

Anaerobic, coupled to DNR, SQR pathway
Anaerobic, coupled to DNR, Sox pathway
Poser et al. (2014) [63]
Elemental sulfurSulfateDesulfocapsa thiozymogenes

(chemolithotroph; "cable bacteria")

Enrichment culture

11.0 to 18.4‰ (28 °C)

12.7 to 17.9‰ (28 °C)

Disproportionation, in the presence of iron scavengersBöttcher et al. (2001) [67]
Desulfocapsa thiozymogenes

(chemolithotroph; "cable bacteria") Enrichment culture

8 to 12 ‰ (28 °C)Disproportionation, attenuated isotope effect due to reoxidation by manganese oxidesBöttcher & Thamdrup (2001) [68]

Fractionation of sulfur isotopes

Aerobic MSO generates depletions in the 34S of sulfate that have been found to be as small as −1.5‰ and as large as -18‰. For most microorganisms and oxidation conditions, only small fractionations accompany either the aerobic or anaerobic oxidation of sulfide, elemental sulfur, thiosulfate and sulfite to elemental sulfur or sulfate. The phototrophic oxidation of sulfide to thiosulfate under anoxic conditions also generates negligible fractionations. Although the change in sulfur isotopes is usually small during MSO, MSO oxidizes reduced forms of sulfur which are usually depleted in 34S compared to seawater sulfate. Therefore large-scale MSO can also significantly change the sulfur isotopes of a reservoir. It has been proposed that the observed global average S-isotope fractionation is around −50‰ instead of the theoretically predicted value -70‰ because of MSO. [69]

In the chemolithotrophs Thiobacillus denitrificans and Sulfurimonas denitrificans, MSO coupled to DNR has the effect of inducing the SQR and Sox pathways, respectively. In both cases, a small fractionation in the 34S of the sulfate, lower than -4.3‰, has been measured. Sulfate depletion in 34S from MSO could be used to trace sulfide oxidation processes in the environment, although it does not allow a discrimination between the SQR and Sox pathways. [63] The depletion produced by MSO coupled to DNR is similar to up to -5‰ depletion estimated for the 34S in the sulfide produced from rDsr. [70] [71] In contrast, disproportionation under anaerobic conditions generates sulfate enriched in 34S up to 9‰ and ~34‰ from sulfide and elemental sulfur, respectively. The isotope effect of disproportionation is however limited by the rates of sulfate reduction and MSO. [72] Just like the fractionation of oxygen isotopes, the larger fractionations in sulfate from the disproportionation of elemental sulfur point to a key step or pathway critical for inducing this large kinetic isotope effect. The table below summarizes the reported fractionations of sulfur isotopes from MSO in different organisms and conditions.

Starting compound (reactant)Intermediate or end compounds
(products)
OrganismAverage 34S fractionation

(product/reactant)

DetailsOxidantReference
SulfideSulfateT. neopolitanus, T. intermedius and T. ferrooxidans (chemolithotrophs)-2 to -5.5‰

(no temperature provided)

Aerobic
pH 5 to 6
Carbon dioxideToran (1986) [73]
Polythionates (SnO62-)
Elemental sulfur
Sulfate
T. concretivorus (chemolithotroph)0.6 to 19‰ (30 °C)
-2.5 to 1.2‰ (30 °C)
-18 to -10.5‰ (30 °C)
AerobicCarbon dioxideKaplan & Rittenberg (1964) [74]
SulfateA. ferrooxidans (chemolithotroph)−1.5‰
−4‰

(no temperature provided)

Aerobic

Anaerobic

Carbon dioxideThurston et al. (2010) [65]
SulfateT. denitrificans (chemolithotroph)
Sulfurimonas denitrificans (chemolithotroph)
−4.3 to −1.3‰ (30 °C)

−2.9 to −1.6‰ (28 °C)

Anaerobic, coupled to DNR, SQR pathway
Anaerobic, coupled to DNR, Sox pathway
Carbon dioxidePoser et al. (2014) [63]
SulfateThiomicrospira sp. strain CVO

(chemolithotroph)

1‰ (no temperature provided)Anaerobic, coupled to DNR, no intermediates in complete oxidation of sulfide to sulfate (potentially only uses Sox pathway)Carbon dioxideHubert et al. (2009) [66]
Elemental sulfurChlorobium thiosulphatophilum
(green sulfur bacteria)
5‰ (no temperature provided)AnaerobicCarbon dioxideKelly et al. (1979) [75]
ThiosulfateOscillatoria sp. (Cyanobacteria)

Calothrix sp. (Cyanobacteria)

0‰ (30 °C)Anaerobic, anoxygenic photosynthesisCarbon dioxideHabicht et al.(1988) [76]
Elemental sulfur

Sulfate

Chromatium vinosum (purple sulfur bacteria)0‰ (30-35 °C)

2‰ (30-35 °C)

Anaerobic, anoxygenic photosynthesisFry et al. (1985) [77]
Elemental sulfur

Sulfate

Ectothiorhodospira shaposhnikovii (purple sulfur bacteria)±5‰ (no temperature provided)Anaerobic, anoxygenic photosynthesisIvanov et al. (1976) [78]
Polythionates (SnO62-)
Elemental sulfur
Sulfate
Chromatium sp. (purple sulfur bacteria)4.9 to 11.2‰ (30 °C)
-10 to -3.6‰ (30 °C)
-2.9 to -0.9‰ (30 °C)
AnaerobicKaplan & Rittenberg (1964) [74]
ThiosulfateSulfateT. intermedius (chemolithotroph)-4.7‰ (no temperature provided)AerobicKelly et al. (1979) [75]
SulfateT. versutus (chemolithotroph)0‰ (28 °C)AerobicFry et al. (1986) [79]
Elemental sulfur + SulfateChromatium vinosum (purple sulfur bacteria)0‰ (30-35 °C)AnaerobicFry et al. (1985) [77]
SulfateDesulfovibrio sulfodismutans

(chemolithotroph)

D. thiozymogenes (chemolithotroph; "cable bacteria")

For both bacteria:

0‰ (30 °C; compared to the sulfonate functional group); 2 to 4‰ (30 °C; compared to the sulfane functional group)

Anaerobic, disproportionation Habicht et al.(1988) [76]
Elemental sulfurSulfateDesulfocapsa thiozymogenes

(chemolithotroph; "cable bacteria")

Enrichment culture

17.4‰ (28 °C)

16.6‰ (28 °C)

Anaerobic, disproportionation, in the presence of iron scavengersBöttcher et al. (2001) [67]
Desulfocapsasulfoexigens

Desulfocapsa thiozymogenes

(chemolithotrophs; "cable bacteria")

Desulfobulbuspropionicus (chemoorganotroph)

Marine enrichments and sediments

16.4‰ (30 °C)

17.4‰ (30 °C)

33.9‰ (35 °C)

17.1 to 20.6‰ (28 °C)

Anaerobic, disproportionation Canfield et al. (1998) [80]
Desulfocapsa thiozymogenes

(chemolithotroph; "cable bacteria")

Enrichment culture

−0.6 to 2.0‰ (28 °C)

−0.2 to 1.1‰ (28 °C)

Anaerobic, disproportionation, attenuated isotope effect due to reoxidation by manganese oxidesBöttcher & Thamdrup (2001) [68]
SulfiteSulfateDesulfovibrio sulfodismutans

(chemolithotroph)

D. thiozymogenes

(chemolithotroph; "cable bacteria")

9 to 12‰ (30 °C)

7 to 9‰ (30 °C)

Anaerobic, disproportionation Habicht et al.(1988) [76]

See also

Related Research Articles

<span class="mw-page-title-main">Sulfur</span> Chemical element, symbol S and atomic number 16

Sulfur (also spelled sulphur in British English) is a chemical element; it has symbol S and atomic number 16. It is abundant, multivalent and nonmetallic. Under normal conditions, sulfur atoms form cyclic octatomic molecules with the chemical formula S8. Elemental sulfur is a bright yellow, crystalline solid at room temperature.

Anaerobic respiration is respiration using electron acceptors other than molecular oxygen (O2). Although oxygen is not the final electron acceptor, the process still uses a respiratory electron transport chain.

<span class="mw-page-title-main">Denitrification</span> Microbially facilitated process

Denitrification is a microbially facilitated process where nitrate (NO3) is reduced and ultimately produces molecular nitrogen (N2) through a series of intermediate gaseous nitrogen oxide products. Facultative anaerobic bacteria perform denitrification as a type of respiration that reduces oxidized forms of nitrogen in response to the oxidation of an electron donor such as organic matter. The preferred nitrogen electron acceptors in order of most to least thermodynamically favorable include nitrate (NO3), nitrite (NO2), nitric oxide (NO), nitrous oxide (N2O) finally resulting in the production of dinitrogen (N2) completing the nitrogen cycle. Denitrifying microbes require a very low oxygen concentration of less than 10%, as well as organic C for energy. Since denitrification can remove NO3, reducing its leaching to groundwater, it can be strategically used to treat sewage or animal residues of high nitrogen content. Denitrification can leak N2O, which is an ozone-depleting substance and a greenhouse gas that can have a considerable influence on global warming.

Methanotrophs are prokaryotes that metabolize methane as their source of carbon and chemical energy. They are bacteria or archaea, can grow aerobically or anaerobically, and require single-carbon compounds to survive.

<span class="mw-page-title-main">Chromatiaceae</span> Family of purple sulfur bacteria

The Chromatiaceae are one of the two families of purple sulfur bacteria, together with the Ectothiorhodospiraceae. They belong to the order Chromatiales of the class Gammaproteobacteria, which is composed by unicellular Gram-negative organisms. Most of the species are photolithoautotrophs and conduct an anoxygenic photosynthesis, but there are also representatives capable of growing under dark and/or microaerobic conditions as either chemolithoautotrophs or chemoorganoheterotrophs.

<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.

In chemistry, disproportionation, sometimes called dismutation, is a redox reaction in which one compound of intermediate oxidation state converts to two compounds, one of higher and one of lower oxidation states. The reverse of disproportionation, such as when a compound in an intermediate oxidation state is formed from precursors of lower and higher oxidation states, is called comproportionation, also known as synproportionation.

<span class="mw-page-title-main">Sulfur cycle</span> Biogeochemical cycle of sulfur

The important sulfur cycle is a biogeochemical cycle in which the sulfur moves between rocks, waterways and living systems. It is important in geology as it affects many minerals and in life because sulfur is an essential element (CHNOPS), being a constituent of many proteins and cofactors, and sulfur compounds can be used as oxidants or reductants in microbial respiration. The global sulfur cycle involves the transformations of sulfur species through different oxidation states, which play an important role in both geological and biological processes. Steps of the sulfur cycle are:

<span class="mw-page-title-main">Sulfur-reducing bacteria</span> Microorganisms able to reduce elemental sulfur to hydrogen sulfide

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.

<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.

Microbial metabolism is the means by which a microbe obtains the energy and nutrients it needs to live and reproduce. Microbes use many different types of metabolic strategies and species can often be differentiated from each other based on metabolic characteristics. The specific metabolic properties of a microbe are the major factors in determining that microbe's ecological niche, and often allow for that microbe to be useful in industrial processes or responsible for biogeochemical cycles.

Anaerobic oxidation of methane (AOM) is a methane-consuming microbial process occurring in anoxic marine and freshwater sediments. AOM is known to occur among mesophiles, but also in psychrophiles, thermophiles, halophiles, acidophiles, and alkophiles. During AOM, methane is oxidized with different terminal electron acceptors such as sulfate, nitrate, nitrite and metals, either alone or in syntrophy with a partner organism.

<span class="mw-page-title-main">Thiosulfate dehydrogenase</span>

Thiosulfate dehydrogenase is an enzyme that catalyzes the chemical reaction:

Sulfur is metabolized by all organisms, from bacteria and archaea to plants and animals. Sulfur can have an oxidation state from -2 to +6 and is reduced or oxidized by a diverse range of organisms. The element is present in proteins, sulfate esters of polysaccharides, steroids, phenols, and sulfur-containing coenzymes.

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.

Desulfobulbus propionicus is a Gram-negative, anaerobic chemoorganotroph. Three separate strains have been identified: 1pr3T, 2pr4, and 3pr10. It is also the first pure culture example of successful disproportionation of elemental sulfur to sulfate and sulfide. Desulfobulbus propionicus has the potential to produce free energy and chemical products.

<i>Acidithiobacillus thiooxidans</i> Species of bacterium

Acidithiobacillus thiooxidans, formerly known as Thiobacillus thiooxidans until its reclassification into the newly designated genus Acidithiobacillus of the Acidithiobacillia subclass of Pseudomonadota, is a Gram-negative, rod-shaped bacterium that uses sulfur as its primary energy source. It is mesophilic, with a temperature optimum of 28 °C. This bacterium is commonly found in soil, sewer pipes, and cave biofilms called snottites. A. thiooxidans is used in the mining technique known as bioleaching, where metals are extracted from their ores through the action of microbes.

The sulfate-methane transition zone (SMTZ) is a zone in oceans, lakes, and rivers typically found below the sediment surface in which sulfate and methane coexist. The formation of a SMTZ is driven by the diffusion of sulfate down the sediment column and the diffusion of methane up the sediments. At the SMTZ, their diffusion profiles meet and sulfate and methane react with one another, which allows the SMTZ to harbor a unique microbial community whose main form of metabolism is anaerobic oxidation of methane (AOM). The presence of AOM marks the transition from dissimilatory sulfate reduction to methanogenesis as the main metabolism utilized by organisms.

<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.

Sulfur isotope biogeochemistry is the study of the distribution of sulfur isotopes in biological and geological materials. In addition to its common isotope, 32S, sulfur has three rare stable isotopes: 34S, 36S, and 33S. The distribution of these isotopes in the environment is controlled by many biochemical and physical processes, including biological metabolisms, mineral formation processes, and atmospheric chemistry. Measuring the abundance of sulfur stable isotopes in natural materials, like bacterial cultures, minerals, or seawater, can reveal information about these processes both in the modern environment and over Earth history.

References

  1. Fry B, Ruf W, Gest H, Hayes JM (1988). "Sulfur isotope effects associated with oxidation of sulfide by O2 in aqueous solution". Isotope Geoscience. 73 (3): 205–10. Bibcode:1988CGIGS..73..205F. doi:10.1016/0168-9622(88)90001-2. PMID   11538336.
  2. Burgin AJ, Hamilton SK (2008). "NO3−-Driven SO42− Production in Freshwater Ecosystems: Implications for N and S Cycling". Ecosystems. 11 (6): 908–922. doi:10.1007/s10021-008-9169-5. S2CID   28390566.
  3. 1 2 3 4 5 6 7 8 9 10 Fike DA, Bradley AS, Leavitt WD (2016). Geomicrobiology of sulfur (Sixth ed.). Ehrlich's Geomicrobiology.
  4. 1 2 3 Luther, George W.; Findlay, Alyssa J.; Macdonald, Daniel J.; Owings, Shannon M.; Hanson, Thomas E.; Beinart, Roxanne A.; Girguis, Peter R. (2011). "Thermodynamics and kinetics of sulfide oxidation by oxygen: a look at inorganically controlled reactions and biologically mediated processes in the environment". Frontiers in Microbiology. 2: 62. doi: 10.3389/fmicb.2011.00062 . ISSN   1664-302X. PMC   3153037 . PMID   21833317.
  5. Hawley, Alyse K.; Brewer, Heather M.; Norbeck, Angela D.; Paša-Tolić, Ljiljana; Hallam, Steven J. (2014-08-05). "Metaproteomics reveals differential modes of metabolic coupling among ubiquitous oxygen minimum zone microbes". Proceedings of the National Academy of Sciences. 111 (31): 11395–11400. Bibcode:2014PNAS..11111395H. doi: 10.1073/pnas.1322132111 . ISSN   0027-8424. PMC   4128106 . PMID   25053816.
  6. Middelburg, Jack J. (2011-12-23). "Chemoautotrophy in the ocean". Geophysical Research Letters. 38 (24): n/a. Bibcode:2011GeoRL..3824604M. doi:10.1029/2011gl049725. hdl: 1874/248832 . ISSN   0094-8276. S2CID   131612365.
  7. Boschker, Henricus T. S.; Vasquez-Cardenas, Diana; Bolhuis, Henk; Moerdijk-Poortvliet, Tanja W. C.; Moodley, Leon (2014-07-08). "Chemoautotrophic Carbon Fixation Rates and Active Bacterial Communities in Intertidal Marine Sediments". PLOS ONE. 9 (7): e101443. Bibcode:2014PLoSO...9j1443B. doi: 10.1371/journal.pone.0101443 . ISSN   1932-6203. PMC   4086895 . PMID   25003508.
  8. Dyksma, Stefan; Bischof, Kerstin; Fuchs, Bernhard M; Hoffmann, Katy; Meier, Dimitri; Meyerdierks, Anke; Pjevac, Petra; Probandt, David; Richter, Michael (2016-02-12). "Ubiquitous Gammaproteobacteria dominate dark carbon fixation in coastal sediments". The ISME Journal. 10 (8): 1939–1953. doi:10.1038/ismej.2015.257. ISSN   1751-7362. PMC   4872838 . PMID   26872043.
  9. Overmann, Jörg; van Gemerden, Hans (2000). "Microbial interactions involving sulfur bacteria: implications for the ecology and evolution of bacterial communities". FEMS Microbiology Reviews. 24 (5): 591–599. doi: 10.1111/j.1574-6976.2000.tb00560.x . ISSN   1574-6976. PMID   11077152.
  10. Jørgensen, Bo Barker; Nelson, Douglas C. (2004). Sulfide oxidation in marine sediments: Geochemistry meets microbiology. In: Sulfur Biogeochemistry - Past and Present. Geological Society of America Special Paper 379. pp. 63–81. doi:10.1130/0-8137-2379-5.63. ISBN   978-0813723792.
  11. Ravenschlag, Katrin; Sahm, Kerstin; Amann, Rudolf (2001-01-01). "Quantitative Molecular Analysis of the Microbial Community in Marine Arctic Sediments (Svalbard)". Applied and Environmental Microbiology. 67 (1): 387–395. Bibcode:2001ApEnM..67..387R. doi:10.1128/AEM.67.1.387-395.2001. ISSN   0099-2240. PMC   92590 . PMID   11133470.
  12. 1 2 3 4 Wasmund, Kenneth; Mußmann, Marc; Loy, Alexander (August 2017). "The life sulfuric: microbial ecology of sulfur cycling in marine sediments". Environmental Microbiology Reports. 9 (4): 323–344. doi:10.1111/1758-2229.12538. ISSN   1758-2229. PMC   5573963 . PMID   28419734.
  13. Winkel, Matthias; Salman-Carvalho, Verena; Woyke, Tanja; Richter, Michael; Schulz-Vogt, Heide N.; Flood, Beverly E.; Bailey, Jake V.; Mußmann, Marc (2016). "Single-cell Sequencing of Thiomargarita Reveals Genomic Flexibility for Adaptation to Dynamic Redox Conditions". Frontiers in Microbiology. 7: 964. doi: 10.3389/fmicb.2016.00964 . ISSN   1664-302X. PMC   4914600 . PMID   27446006.
  14. Dubilier, Nicole; Bergin, Claudia; Lott, Christian (2008). "Symbiotic diversity in marine animals: the art of harnessing chemosynthesis". Nature Reviews. Microbiology. 6 (10): 725–740. doi:10.1038/nrmicro1992. ISSN   1740-1534. PMID   18794911. S2CID   3622420.
  15. Risgaard-Petersen, Nils; Revil, André; Meister, Patrick; Nielsen, Lars Peter (2012). "Sulfur, iron-, and calcium cycling associated with natural electric currents running through marine sediment". Geochimica et Cosmochimica Acta. 92: 1–13. Bibcode:2012GeCoA..92....1R. doi:10.1016/j.gca.2012.05.036. ISSN   0016-7037.
  16. Nielsen, Lars Peter; Risgaard-Petersen, Nils; Fossing, Henrik; Christensen, Peter Bondo; Sayama, Mikio (2010). "Electric currents couple spatially separated biogeochemical processes in marine sediment". Nature. 463 (7284): 1071–1074. Bibcode:2010Natur.463.1071N. doi:10.1038/nature08790. ISSN   0028-0836. PMID   20182510. S2CID   205219761.
  17. Seitaj, Dorina; Schauer, Regina; Sulu-Gambari, Fatimah; Hidalgo-Martinez, Silvia; Malkin, Sairah Y.; Burdorf, Laurine D. W.; Slomp, Caroline P.; Meysman, Filip J. R. (2015-10-27). "Cable bacteria generate a firewall against euxinia in seasonally hypoxic basins". Proceedings of the National Academy of Sciences. 112 (43): 13278–13283. Bibcode:2015PNAS..11213278S. doi: 10.1073/pnas.1510152112 . ISSN   0027-8424. PMC   4629370 . PMID   26446670.
  18. Malkin, Sairah Y; Rao, Alexandra MF; Seitaj, Dorina; Vasquez-Cardenas, Diana; Zetsche, Eva-Maria; Hidalgo-Martinez, Silvia; Boschker, Henricus TS; Meysman, Filip JR (2014-03-27). "Natural occurrence of microbial sulphur oxidation by long-range electron transport in the seafloor". The ISME Journal. 8 (9): 1843–1854. doi:10.1038/ismej.2014.41. ISSN   1751-7362. PMC   4139731 . PMID   24671086.
  19. Fuchs T, Huber H, Burggraf S, Stetter KO (1996). "16S rDNA-based Phylogeny of the Archaeal Order Sulfolobales and Reclassification of Desulfurolobus ambivalens as Acidianus ambivalens comb. nov". Systematic and Applied Microbiology. 19 (1): 56–60. doi:10.1016/s0723-2020(96)80009-9. ISSN   0723-2020.
  20. Stetter KO (2002). "Hyperthermophilic Microorganisms" (PDF). Astrobiology. Springer, Berlin, Heidelberg. pp. 169–184. doi:10.1007/978-3-642-59381-9_12. ISBN   9783642639579.
  21. 1 2 Kelly DP, Wood AP (March 2000). "Reclassification of some species of Thiobacillus to the newly designated genera Acidithiobacillus gen. nov., Halothiobacillus gen. nov. and Thermithiobacillus gen. nov". International Journal of Systematic and Evolutionary Microbiology. 50 Pt 2 (2): 511–16. doi: 10.1099/00207713-50-2-511 . PMID   10758854.
  22. 1 2 3 Friedrich CG, Mitrenga G (1981). "Oxidation of thiosulfate by Paracoccus denitrificans and other hydrogen bacteria". FEMS Microbiology Letters. 10 (2): 209–212. doi: 10.1111/j.1574-6968.1981.tb06239.x .
  23. Huber R, Eder W (2006). The Prokaryotes. Springer, New York, NY. pp. 925–938. doi:10.1007/0-387-30747-8_39. ISBN   9780387254975.
  24. Aragno M (1992). "The aerobic, chemolithoautotrophic, thermophilic bacteria". In Kristjansson JK (ed.). Thermophilic Bacteria. Boca Raton, Fla.: CRC Press. pp. 77–104.
  25. Kelly DP, Smith NA (1990). Advances in Microbial Ecology. Advances in Microbial Ecology. Springer, Boston, MA. pp. 345–385. doi:10.1007/978-1-4684-7612-5_9. ISBN   9781468476149.
  26. Kelly DP, McDonald IR, Wood AP (September 2000). "Proposal for the reclassification of Thiobacillus novellus as Starkeya novella gen. nov., comb. nov., in the alpha-subclass of the Proteobacteria". International Journal of Systematic and Evolutionary Microbiology. 50 Pt 5 (5): 1797–802. doi: 10.1099/00207713-50-5-1797 . PMID   11034489.
  27. Trojan, Daniela; Schreiber, Lars; Bjerg, Jesper T.; Bøggild, Andreas; Yang, Tingting; Kjeldsen, Kasper U.; Schramm, Andreas (2016). "A taxonomic framework for cable bacteria and proposal of the candidate genera Electrothrix and Electronema". Systematic and Applied Microbiology. 39 (5): 297–306. doi:10.1016/j.syapm.2016.05.006. ISSN   0723-2020. PMC   4958695 . PMID   27324572.
  28. Imhoff JF, Süling J, Petri R (October 1998). "Phylogenetic relationships among the Chromatiaceae, their taxonomic reclassification and description of the new genera Allochromatium, Halochromatium, Isochromatium, Marichromatium, Thiococcus, Thiohalocapsa and Thermochromatium". International Journal of Systematic Bacteriology. 48 Pt 4 (4): 1129–43. doi: 10.1099/00207713-48-4-1129 . PMID   9828415.
  29. Imhoff JF, Suling J, Petri R (1 October 1998). "Phylogenetic relationships among the Chromatiaceae, their taxonomic reclassification and description of the new genera Allochromatium, Halochromatium, Isochromatium, Marichromatium, Thiococcus, Thiohalocapsa and Thermochromatium". International Journal of Systematic Bacteriology. 48 (4): 1129–1143. doi: 10.1099/00207713-48-4-1129 . PMID   9828415.
  30. Brune DC (July 1989). "Sulfur oxidation by phototrophic bacteria". Biochimica et Biophysica Acta (BBA) - Bioenergetics. 975 (2): 189–221. doi:10.1016/S0005-2728(89)80251-8. PMID   2663079.
  31. Cohen Y, Padan E, Shilo M (September 1975). "Facultative anoxygenic photosynthesis in the cyanobacterium Oscillatoria limnetica". Journal of Bacteriology. 123 (3): 855–61. doi:10.1128/jb.123.3.855-861.1975. PMC   235807 . PMID   808537.
  32. Garlick S, Oren A, Padan E (February 1977). "Occurrence of facultative anoxygenic photosynthesis among filamentous and unicellular cyanobacteria". Journal of Bacteriology. 129 (2): 623–29. doi:10.1128/jb.129.2.623-629.1977. PMC   234991 . PMID   402355.
  33. Beller HR, Chain PS, Letain TE, Chakicherla A, Larimer FW, Richardson PM, Coleman MA, Wood AP, Kelly DP (February 2006). "The genome sequence of the obligately chemolithoautotrophic, facultatively anaerobic bacterium Thiobacillus denitrificans". Journal of Bacteriology. 188 (4): 1473–88. doi:10.1128/JB.188.4.1473-1488.2006. PMC   1367237 . PMID   16452431.
  34. Turchyn AV, Brüchert V, Lyons TW, Engel GS, Balci N, Schrag DP, Brunner B (2010). "Kinetic oxygen isotope effects during dissimilatory sulfate reduction: A combined theoretical and experimental approach". Geochimica et Cosmochimica Acta. 74 (7): 2011–2024. Bibcode:2010GeCoA..74.2011T. doi:10.1016/j.gca.2010.01.004. ISSN   0016-7037.
  35. Yamanaka, T.; Fukumori, Y.; Okunuki, K. (1979). "Preparation of subunits of flavocytochromes c derived from Chlorobium limicola f. thiosulfatophilum and Chromatium vinosum". Analytical Biochemistry. 95 (1): 209–213. doi:10.1016/0003-2697(79)90207-0. ISSN   0003-2697. PMID   227287.
  36. Sievert SM, Scott KM, Klotz MG, Chain PS, Hauser LJ, Hemp J, Hügler M, Land M, Lapidus A, Larimer FW, Lucas S, Malfatti SA, Meyer F, Paulsen IT, Ren Q, Simon J (February 2008). "Genome of the epsilonproteobacterial chemolithoautotroph Sulfurimonas denitrificans". Applied and Environmental Microbiology. 74 (4): 1145–56. Bibcode:2008ApEnM..74.1145S. doi:10.1128/AEM.01844-07. PMC   2258580 . PMID   18065616.
  37. 1 2 Kelly DP (1989). "Physiology and biochemistry of unicellular sulfur bacteria". In Schlegel HG; Bowien B (eds.). Autotrophic Bacteria. FEMS Symposium. Springer-Verlag. pp. 193–217.
  38. 1 2 Kelly DP, Shergill JK, Lu WP & Wood AP (1997). "Oxidative metabolism of inorganic sulfur compounds by bacteria". Antonie van Leeuwenhoek. 71 (1–2): 95–107. doi:10.1023/A:1000135707181. PMID   9049021. S2CID   2057300.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  39. 1 2 3 Friedrich CG, Rother D, Bardischewsky F, Quentmeier A, Fischer J (July 2001). "Oxidation of reduced inorganic sulfur compounds by bacteria: emergence of a common mechanism?". Applied and Environmental Microbiology. 67 (7): 2873–82. Bibcode:2001ApEnM..67.2873F. doi:10.1128/AEM.67.7.2873-2882.2001. PMC   92956 . PMID   11425697.
  40. Grimm F, Franz B, Dahl C (2008). "Thiosulfate and sulfur oxidation in purple sulfur bacteria.". In Friedrich C, Dahl C (eds.). Microbial Sulfur Metabolism. Berlin Heidelberg: Springer-Verlag GmbH. pp. 101–116.
  41. Kappler, U.; Dahl, C. (2001-09-11). "Enzymology and molecular biology of prokaryotic sulfite oxidation". FEMS Microbiology Letters. 203 (1): 1–9. doi: 10.1111/j.1574-6968.2001.tb10813.x . ISSN   0378-1097. PMID   11557133.
  42. Ghosh, Wriddhiman; Dam, Bomba (2009). "Biochemistry and molecular biology of lithotrophic sulfur oxidation by taxonomically and ecologically diverse bacteria and archaea". FEMS Microbiology Reviews. 33 (6): 999–1043. doi: 10.1111/j.1574-6976.2009.00187.x . ISSN   1574-6976. PMID   19645821.
  43. Dahl, C., Rákhely, G., Pott-Sperling, A. S., Fodor, B., Takács, M., Tóth, A., Kraeling, M., Győrfi, K., Kovács, A., Tusz, J. & Kovács, K. L. (1999). "Genes involved in hydrogen and sulfur metabolism in phototrophic sulfur bacteria". FEMS Microbiology Letters. 180 (2): 317–324. doi: 10.1111/j.1574-6968.1999.tb08812.x . ISSN   0378-1097. PMID   10556728.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  44. Rivett MO, Buss SR, Morgan P, Smith JW, Bemment CD (October 2008). "Nitrate attenuation in groundwater: a review of biogeochemical controlling processes". Water Research. 42 (16): 4215–32. Bibcode:2008WatRe..42.4215R. doi:10.1016/j.watres.2008.07.020. PMID   18721996.
  45. Wit R, Gemerden H (1987). "Oxidation of sulfide to thiosulfate by Microcoleus chtonoplastes". FEMS Microbiology Letters. 45 (1): 7–13. doi: 10.1111/j.1574-6968.1987.tb02332.x .
  46. Rabenstein A, Rethmeier J, Fischer U (2014). "Sulphite as Intermediate Sulphur Compound in Anaerobic Sulphide Oxidation to Thiosulphate by Marine Cyanobacteria". Zeitschrift für Naturforschung C. 50 (11–12): 769–774. doi: 10.1515/znc-1995-11-1206 .
  47. Wood P (1988). "Chemolithotrophy". In Anthony C (ed.). Bacterial Energy Transduction. London, UK: Academic Press. pp. 183–230.
  48. Roy AB, Trudinger PA (1970). The biochemistry of inorganic compounds of sulphur. Cambridge University Press.
  49. Zimmermann P, Laska S, Kletzin A (August 1999). "Two modes of sulfite oxidation in the extremely thermophilic and acidophilic archaeon acidianus ambivalens". Archives of Microbiology. 172 (2): 76–82. doi:10.1007/s002030050743. PMID   10415168. S2CID   9216478.
  50. Aminuddin M (November 1980). "Substrate level versus oxidative phosphorylation in the generation of ATP in Thiobacillus denitrificans". Archives of Microbiology. 128 (1): 19–25. doi:10.1007/BF00422300. PMID   7458535. S2CID   13042589.
  51. Aminuddin M, Nicholas DJ (1974). "Electron Transfer during Sulphide and Sulphite Oxidation in Thiobacillus denitrificans". Microbiology. 82 (1): 115–123. doi: 10.1099/00221287-82-1-115 .
  52. Aminuddin M, Nicholas DJ (October 1973). "Sulphide oxidation linked to the reduction of nitrate and nitrite in Thiobacillus denitrificans". Biochimica et Biophysica Acta (BBA) - Bioenergetics. 325 (1): 81–93. doi:10.1016/0005-2728(73)90153-9. PMID   4770733.
  53. Gevertz D, Telang AJ, Voordouw G, Jenneman GE (June 2000). "Isolation and characterization of strains CVO and FWKO B, two novel nitrate-reducing, sulfide-oxidizing bacteria isolated from oil field brine". Applied and Environmental Microbiology. 66 (6): 2491–501. Bibcode:2000ApEnM..66.2491G. doi:10.1128/AEM.66.6.2491-2501.2000. PMC   110567 . PMID   10831429.
  54. Burton SD, Morita RY (December 1964). "Effect of Catalase and Cultural Conditions on Growth of Beggiatoa". Journal of Bacteriology. 88 (6): 1755–61. doi:10.1128/jb.88.6.1755-1761.1964. PMC   277482 . PMID   14240966.
  55. Skerman VB, Dementjeva G, Carey BJ (April 1957). "Intracellular deposition of sulfur by Sphaerotilus natans". Journal of Bacteriology. 73 (4): 504–12. doi:10.1128/jb.73.4.504-512.1957. PMC   314609 . PMID   13428683.
  56. Skerman VB, Dementjev G, Skyring GW (April 1957). "Deposition of sulphur from hydrogen sulphide by bacteria and yeast". Nature. 179 (4562): 742. Bibcode:1957Natur.179..742S. doi: 10.1038/179742a0 . PMID   13418779. S2CID   4284363.
  57. Belousova EV, Chernousova EI, Dubinina GA, Turova TP, Grabovich MI (2013). "[Detection and analysis of sulfur metabolism genes in Sphaerotilus natans subsp. sulfidivorans representatives]". Mikrobiologiia. 82 (5): 579–87. PMID   25509396.
  58. Suzuki I, Silver M (July 1966). "The initial product and properties of the sulfur-oxidizing enzyme of thiobacilli". Biochimica et Biophysica Acta (BBA) - Enzymology and Biological Oxidation. 122 (1): 22–33. doi:10.1016/0926-6593(66)90088-9. PMID   5968172.
  59. Brock TD, Gustafson J (October 1976). "Ferric iron reduction by sulfur- and iron-oxidizing bacteria". Applied and Environmental Microbiology. 32 (4): 567–71. Bibcode:1976ApEnM..32..567B. doi:10.1128/aem.32.4.567-571.1976. PMC   170307 . PMID   825043.
  60. Lu W-P. (1986). "A periplasmic location for the bisulfiteoxidizing multienzyme system from Thiobacillus versutus". FEMS Microbiol Lett. 34 (3): 313–317. doi: 10.1111/j.1574-6968.1986.tb01428.x .
  61. Pronk, J (1990). "Oxidation of reduced inorganic sulphur compounds by acidophilic thiobacilli". FEMS Microbiology Letters. 75 (2–3): 293–306. doi: 10.1111/j.1574-6968.1990.tb04103.x . ISSN   0378-1097.
  62. Knöller K, Vogt C, Feisthauer S, Weise SM, Weiss H, Richnow HH (November 2008). "Sulfur cycling and biodegradation in contaminated aquifers: insights from stable isotope investigations". Environmental Science & Technology. 42 (21): 7807–12. Bibcode:2008EnST...42.7807V. doi:10.1021/es800331p. PMID   19031864.
  63. 1 2 3 4 5 Poser A, Vogt C, Knöller K, Ahlheim J, Weiss H, Kleinsteuber S, Richnow HH (August 2014). "Stable sulfur and oxygen isotope fractionation of anoxic sulfide oxidation by two different enzymatic pathways". Environmental Science & Technology. 48 (16): 9094–102. Bibcode:2014EnST...48.9094P. doi:10.1021/es404808r. PMID   25003498.
  64. Taylor BE, Wheeler MC, Nordstrom DK (1984). "Stable isotope geochemistry of acid mine drainage: Experimental oxidation of pyrite". Geochimica et Cosmochimica Acta. 48 (12): 2669–2678. Bibcode:1984GeCoA..48.2669T. doi:10.1016/0016-7037(84)90315-6. ISSN   0016-7037.
  65. 1 2 Thurston RS, Mandernack KW, Shanks WC (2010). "Laboratory chalcopyrite oxidation by Acidithiobacillus ferrooxidans: Oxygen and sulfur isotope fractionation". Chemical Geology. 269 (3–4): 252–261. Bibcode:2010ChGeo.269..252T. doi:10.1016/j.chemgeo.2009.10.001.
  66. 1 2 Hubert C, Voordouw G, Mayer B (2009). "Elucidating microbial processes in nitrate- and sulfate-reducing systems using sulfur and oxygen isotope ratios: The example of oil reservoir souring control". Geochimica et Cosmochimica Acta. 73 (13): 3864–3879. Bibcode:2009GeCoA..73.3864H. doi:10.1016/j.gca.2009.03.025.
  67. 1 2 Böttcher ME, Thamdrup B, Vennemann TW (2001). "Oxygen and sulfur isotope fractionation during anaerobic bacterial disproportionation of elemental sulfur". Geochimica et Cosmochimica Acta. 65 (10): 1601–1609. Bibcode:2001GeCoA..65.1601B. doi:10.1016/s0016-7037(00)00628-1.
  68. 1 2 Böttcher ME, Thamdrup B (2001). "Anaerobic sulfide oxidation and stable isotope fractionation associated with bacterial sulfur disproportionation in the presence of MnO2". Geochimica et Cosmochimica Acta. 65 (10): 1573–1581. Bibcode:2001GeCoA..65.1573B. doi:10.1016/s0016-7037(00)00622-0.
  69. Tsang, Man-Yin; Wortmann, Ulrich G. (2022-07-06). "Sulfur isotope fractionation derived from reaction-transport modelling in the Eastern Equatorial Pacific". Journal of the Geological Society. 179 (5). Bibcode:2022JGSoc.179...68T. doi:10.1144/jgs2021-068. ISSN   0016-7649. S2CID   248647580.
  70. Brunner B, Bernasconi SM, Kleikemper J, Schroth MH (2005). "A model for oxygen and sulfur isotope fractionation in sulfate during bacterial sulfate reduction processes". Geochimica et Cosmochimica Acta. 69 (20): 4773–4785. Bibcode:2005GeCoA..69.4773B. doi:10.1016/j.gca.2005.04.017. ISSN   0016-7037.
  71. Brunner B, Bernasconi SM (2005). "A revised isotope fractionation model for dissimilatory sulfate reduction in sulfate reducing bacteria". Geochimica et Cosmochimica Acta. 69 (20): 4759–4771. Bibcode:2005GeCoA..69.4759B. doi:10.1016/j.gca.2005.04.015. ISSN   0016-7037.
  72. Tsang, Man-Yin; Böttcher, Michael Ernst; Wortmann, Ulrich Georg (2023-08-20). "Estimating the effect of elemental sulfur disproportionation on the sulfur-isotope signatures in sediments". Chemical Geology. 632: 121533. doi:10.1016/j.chemgeo.2023.121533. ISSN   0009-2541. S2CID   258600480.
  73. Toran L. (1986) Sulfate contamination in groundwater near an abandoned mine: Hydrogeochemical modeling, microbiology and isotope geochemistry. PhD. dissertation, Univ. of Wisconsin.
  74. 1 2 Kaplan IR, Rittenberg SC (1964). "Microbiological Fractionation of Sulphur Isotopes". Journal of General Microbiology. 34 (2): 195–212. doi: 10.1099/00221287-34-2-195 . PMID   14135528.
  75. 1 2 Kelly DP, Chambers LA, Rafter TA (1979). "Unpublished results. In: Microbiological fractionation of stable sulfur isotopes: a review and critique". Geomicrobiology Journal. 1 (3): 249–293. doi:10.1080/01490457909377735.
  76. 1 2 3 Habicht KS, Canfield DE, Rethmeier J (1998). "Sulfur isotope fractionation during bacterial reduction and disproportionation of thiosulfate and sulfite". Geochimica et Cosmochimica Acta. 62 (15): 2585–2595. Bibcode:1998GeCoA..62.2585H. doi:10.1016/s0016-7037(98)00167-7.
  77. 1 2 Fry B, Gest H, Hayes J (1985). "Isotope effects associated with the anaerobic oxidation of sulfite and thiosulfate by the photosynthetic bacterium,Chromatium vinosum". FEMS Microbiology Letters. 27 (2): 227–232. doi: 10.1111/j.1574-6968.1985.tb00672.x . ISSN   0378-1097. PMID   11540842.
  78. Ivanov MV, Gogotova GI, Matrosov AG, Ziakun AM (1976). "[Fractionation of sulfur isotopes by phototrophic sulfur bacterium Ectothiorhodospira shaposhnikovii]". Mikrobiologiia. 45 (5): 757–62. PMID   1004261.
  79. Fry B, Cox J, Gest H, Hayes JM (January 1986). "Discrimination between 34S and 32S during bacterial metabolism of inorganic sulfur compounds". Journal of Bacteriology. 165 (1): 328–30. doi:10.1128/jb.165.1.328-330.1986. PMC   214413 . PMID   3941049.
  80. Canfield DE, Thamdrup B, Fleischer S (1998). "Isotope fractionation and sulfur metabolism by pure and enrichment cultures of elemental sulfur-disproportionating bacteria". Limnology and Oceanography. 43 (2): 253–264. Bibcode:1998LimOc..43..253C. doi: 10.4319/lo.1998.43.2.0253 . hdl: 21.11116/0000-0005-1BA7-1 .