Iron-oxidizing bacteria

Last updated
Iron-oxidizing bacteria in surface water Iron bacteria in runoff.JPG
Iron-oxidizing bacteria in surface water

Iron-oxidizing bacteria (or iron 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. [1]

Contents

When de-oxygenated water reaches a source of oxygen, iron bacteria convert dissolved iron into an insoluble reddish-brown gelatinous slime that discolors stream beds and can stain plumbing fixtures, clothing, or utensils washed with the water carrying it. [2]

Organic material dissolved in water is often the underlying cause of an iron-oxidizing bacteria population. Groundwater may be naturally de-oxygenated by decaying vegetation in swamps. Useful mineral deposits of bog iron ore have formed where groundwater has historically emerged and been exposed to atmospheric oxygen. [3] Anthropogenic hazards like landfill leachate, septic drain fields, or leakage of light petroleum fuels like gasoline are other possible sources of organic materials allowing soil microbes to de-oxygenate groundwater. [4]

A similar reaction may form black deposits of manganese dioxide from dissolved manganese but is less common because of the relative abundance of iron (5.4%) in comparison to manganese (0.1%) in average soils. [5] The sulfurous smell of rot or decay sometimes associated with iron-oxidizing bacteria results from the enzymatic conversion of soil sulfates to volatile hydrogen sulfide as an alternative source of oxygen in anaerobic water. [6]

Iron is a very important chemical element required by living organisms to carry out numerous metabolic reactions such as the formation of proteins involved in biochemical reactions. Examples of these proteins include iron–sulfur proteins, hemoglobin, and coordination complexes. Iron has a widespread distribution globally and is considered one of the most abundant elements in the Earth's crust, soil, and sediments. Iron is a trace element in marine environments. [7] Its role as the electron donor of some chemolithotrophs is probably very ancient. [8]

Metabolism

The anoxygenic phototrophic iron oxidation was the first anaerobic metabolism to be described within the iron anaerobic oxidation metabolism. The photoferrotrophic bacteria use Fe2+ as electron donor and the energy from light to assimilate CO2 into biomass through the Calvin Benson-Bassam cycle (or rTCA cycle) in a neutrophilic environment (pH 5.5-7.2), producing Fe3+ oxides as a waste product that precipitates as a mineral, according to the following stoichiometry (4 mM of Fe(II) can yield 1 mM of CH2O):

HCO3 + 4Fe(II) + 10H2O → [CH2O] + 4Fe(OH)3 + 7H+ (∆G° > 0) [7] [9]

Nevertheless, some bacteria do not use the photoautotrophic Fe(II) oxidation metabolism for growth purposes. [10] Instead, it has been suggested that these groups are sensitive to Fe(II) and therefore oxidize Fe(II) into more insoluble Fe(III) oxide to reduce its toxicity, enabling them to grow in the presence of Fe(II). [10] On the other hand, based on experiments with R. capsulatus SB1003 (photoheterotrophic), it has been demonstrated that the oxidation of Fe(II) might be the mechanisms whereby the bacteria is enabled to access organic carbon sources (acetate, succinate) whose use depends on Fe(II) oxidation [11] Nonetheless, many iron-oxidizing bacteria can use other compounds as electron donors in addition to Fe(II), or even perform dissimilatory Fe(III) reduction as the Geobacter metallireducens . [10]

The dependence of photoferrotrophics on light as a crucial resource [12] [9] [13] can take the bacteria to a cumbersome situation, where due to their requirement for anoxic lighted regions (near the surface) [9] they could be faced with competition by abiotic reactions due to the presence of molecular oxygen. To avoid this problem, they tolerate microaerophilic surface conditions or perform the photoferrotrophic Fe(II) oxidation deeper in the sediment/water column, with low light availability. [9]

Light penetration can limit the Fe(II) oxidation in the water column. [12] However, nitrate dependent microbial Fe(II) oxidation is a light independent metabolism that has been shown to support microbial growth in various freshwater and marine sediments (paddy soil, stream, brackish lagoon, hydrothermal, deep-sea sediments) and later on demonstrated as a pronounced metabolism within the water column at the oxygen minimum zone. [14] [13] Microbes that perform this metabolism are successful in neutrophilic or alcaline environments, due to the high difference in between the redox potential of the couples Fe2+/Fe3+ and NO3/NO2 (+200 mV and +770 mV, respectively) releasing a lot of free energy when compared to other iron oxidation metabolisms. [10] [15]

2Fe2+ + NO3 + 5H2O → 2Fe(OH)3 + NO2 + 4H+ (∆G°=-103.5 kJ/mol)

The microbial oxidation of ferrous iron coupled to denitrification (with nitrite or dinitrogen gas being the final product) [7] can be autotrophic using inorganic carbon or organic co-substrates (acetate, butyrate, pyruvate, ethanol) performing heterotrophic growth in the absence of inorganic carbon. [10] [15] It has been suggested that the heterotrophic nitrate-dependent ferrous iron oxidation using organic carbon might be the most favorable process. [16] This metabolism might be very important for carrying out an important step in the biogeochemical cycle within the OMZ. [17]

Types

Despite being phylogenetically diverse, the microbial ferrous iron oxidation metabolic strategy (found in Archaea and Bacteria) is present in 7 phyla, being highly pronounced in the phylum Pseudomonadota (formerly Proteobacteria), particularly the Alpha, Beta, Gamma, and Zetaproteobacteria classes, [10] [18] and among the Archaea domain in the "Euryarchaeota" and Thermoproteota phyla, as well as in Actinomycetota, Bacillota, Chlorobiota, and Nitrospirota phyla. [18]

There are very well-studied iron-oxidizing bacterial species such as Thiobacillus ferrooxidans , and Leptospirillum ferrooxidans, and some like Gallionella ferruginea and Mariprofundis ferrooxydans are able to produce a particular extracellular stalk-ribbon structure rich in iron, known as a typical biosignature of microbial iron oxidation. These structures can be easily detected in a sample of water, indicating the presence iron-oxidizing bacteria. This biosignature has been a tool to understand the importance of iron metabolism in the Earth's past. [19]

Habitat

Iron-oxidizing bacteria colonize the transition zone where de-oxygenated water from an anaerobic environment flows into an aerobic environment. Groundwater containing dissolved organic material may be de-oxygenated by microorganisms feeding on that dissolved organic material. In aerobic conditions, pH variation plays an important role in driving the oxidation reaction of Fe2+/Fe3+. [7] [13] At neutrophilic pHs (hydrothermal vents, deep ocean basalts, groundwater iron seeps) the oxidation of iron by microorganisms is highly competitive with the rapid abiotic reaction occurring in <1 min. [20] Therefore, the microbial community has to inhabit microaerophilic regions where the low oxygen concentration allows the cell to oxidize Fe(II) and produce energy to grow. [21] [22] However, under acidic conditions, where ferrous iron is more soluble and stable even in the presence of oxygen, only biological processes are responsible for the oxidation of iron, [9] thus making ferrous iron oxidation the major metabolic strategy in iron-rich acidic environments. [18] [7]

In the marine environment, the most well-known class of iron oxidizing-bacteria is zetaproteobacteria, [23] which are major players in marine ecosystems. Being generally microaerophilic they are adapted to live in transition zones where the oxic and anoxic waters mix. [21] The zetaproteobacteria are present in different Fe(II)-rich habitats, found in deep ocean sites associated with hydrothermal activity and in coastal and terrestrial habitats, and have been reported in the surface of shallow sediments, beach aquifer, and surface water.

Mariprofundus ferrooxydans is one of the most common and well-studied species of zetaproteobacteria. It was first isolated from the Kamaʻehuakanaloa Seamount (formerly Loihi) vent field, near Hawaii [18] at a depth between 1100 and 1325 meters, on the summit of this shield volcano. Vents can be found ranging from slightly above ambient (10 °C) to high temperature (167 °C). The vent waters are rich in CO2, Fe(II) and Mn. [24] Large, heavily encrusted mats with a gelatinous texture are created by iron-oxidizing bacteria as a by-product (iron-oxyhydroxide precipitation), and can be present around the vent orifices. The vents present at Kamaʻehuakanaloa seamount can be categorized into two types based on concentration and temperature of flow. Those with a focused and high-temperature flow (above 50 °C) can be expected to show higher flow rates as well. These vents are characterized by flocculent mats aggregated around the vent orifices. Mat depth at focused, high-temperature vents averages in the tens of centimeters, but can vary. In contrast, vents with cooler (10-30 °C) and diffuse flow can create mats up to one meter thick. These mats may cover hundreds of square meters of sea floor. [18] Either type of mat can be colonized by other bacterial communities, which can change the chemical composition and the flow of the local waters. [25]

Impact on early life on Earth

Unlike most lithotrophic metabolisms, the oxidation of Fe2+ to Fe3+ yields very little energy to the cell (∆G° = 29 kJ/mol and ∆G° = -90 kJ/mol in acidic and neutral environments, respectively) compared to other chemolithotrophic metabolisms. [18] Therefore, the cell must oxidize large amounts of Fe2+ to fulfill its metabolic requirements while contributing to the mineralization process (through the excretion of twisted stalks). [7] [26] The aerobic iron-oxidizing bacterial metabolism is thought to have made a remarkable contribution to the formation of the largest iron deposit (banded iron formation (BIF)) due to the advent of oxygen in the atmosphere 2.7 billion years ago (produced by cyanobacteria). [13]

However, with the discovery of Fe(II) oxidation carried out under anoxic conditions in the late 1990s [16] using light as an energy source or chemolithotrophically, using a different terminal electron acceptor (mostly NO3), [9] the suggestion arose that anoxic Fe2+ metabolism may pre-date aerobic Fe2+ oxidation and that the age of the BIF pre-dates oxygenic photosynthesis. [7] This suggests that microbial anoxic phototrophic and anaerobic chemolithotrophic metabolism may have been present on the ancient earth, and together with Fe(III) reducers, they may have been responsible for the BIF in the Precambrian eon. [9]

Impact of climate change

In open ocean systems full of dissolved iron, iron-oxidizing bacterial metabolism is ubiquitous and influences the iron cycle. Nowadays, this biochemical cycle is undergoing modifications due to pollution and climate change; nonetheless, the normal distribution of ferrous iron in the ocean could be affected by global warming under the following conditions: acidification, shifting of ocean currents, and ocean water and groundwater hypoxia trend. [20]

These are all consequences of the substantial increase of CO2 emissions into the atmosphere from anthropogenic sources. Currently the concentration of carbon dioxide in the atmosphere is around 420 ppm (120 ppm more than 20 million years ago), and about a quarter of the total CO2 emission enters the oceans (2.2 pg C year−1). Reacting with seawater it produces bicarbonate ion (HCO3) and thus the ocean acidity increases. Furthermore, the temperature of the ocean has increased by almost one degree (0.74 °C) causing the melting of big quantities of glaciers contributing to the sea-level rise. This lowers the O2 solubility by inhibiting the oxygen exchange between surface waters, where O2 is very abundant, and anoxic deep waters. [27] [28]

All these changes in the marine parameters (temperature, acidity, and oxygenation) impact the iron biogeochemical cycle and could have several and critical implications on ferrous iron oxidizing microbes; hypoxic and acid conditions could improve primary productivity in the superficial and coastal waters because that would increase the availability of ferrous iron Fe(II) for microbial iron oxidation. Still, at the same time, this scenario could also disrupt the cascade effect to the sediment in deep water and cause the death of benthonic animals. Moreover it is very important to consider that iron and phosphate cycles are strictly interconnected and balanced, so that a small change in the first could have substantial consequences on the second. [29]

Influence on water infrastructure

A burn in Scotland with iron-oxidizing bacteria Iron bacteria burn.JPG
A burn in Scotland with iron-oxidizing bacteria

Iron-oxidizing bacteria can pose an issue for the management of water-supply wells, as they can produce insoluble ferric oxide, which appears as brown gelatinous slime that will stain plumbing fixtures, and clothing or utensils washed with the water carrying it.

The dramatic effects of iron bacteria are seen in surface waters as brown slimy masses on stream bottoms and lakeshores or as an oily sheen upon the water. More serious problems occur when bacteria build up in well systems. Iron bacteria in wells do not cause health problems, but they can reduce well yields by clogging screens and pipes.

Treatment techniques that may successfully remove or reduce iron bacteria include physical removal, pasteurization, and chemical treatment. Treatment of heavily infected wells may be difficult, expensive, and only partially successful. [30] Recent application of ultrasonic devices that destroy and prevent the formation of biofilm in wells has been proven to prevent iron bacteria infection and the associated clogging very successfully. [31] [32]

Physical removal is typically done as a first step. Small diameter pipes are sometimes cleaned with a wire brush, while larger lines can be scrubbed and flushed clean with a sewer jetter. [33] The pumping equipment in the well must also be removed and cleaned. [34]

Iron filters have been used to treat iron bacteria. Iron filters are similar in appearance and size to conventional water softeners but contain beds of media that have mild oxidizing power. As the iron-bearing water is passed through the bed, any soluble ferrous iron is converted to the insoluble ferric state and then filtered from the water. Any previously precipitated iron is removed by simple mechanical filtration. Several different filter media may be used in these iron filters, including manganese greensand, Birm, MTM, multi-media, sand, and other synthetic materials. In most cases, the higher oxides of manganese produce the desired oxidizing action. Iron filters do have limitations; since the oxidizing action is relatively mild, it will not work well when organic matter, either combined with the iron or completely separate, is present in the water. As a result, the iron bacteria will not be killed. Extremely high iron concentrations may require inconvenient frequent backwashing and/or regeneration. Finally, iron filter media requires high flow rates for proper backwashing, and such water flows are not always available.

Wildfires may release iron-containing compounds from the soil into small wildland streams and cause a rapid but usually temporary proliferation of iron-oxidizing bacteria complete with orange coloration, gelatinous mats, and sulfurous odors. Higher quality personal filters may be used to remove bacteria, odor and restore water clarity.

See also

Related Research Articles

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">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.

<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.

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. These bacteria form colorless filaments composed of cells that can be up to 200 µm in diameter, and are one of the largest prokaryotes on Earth. Beggiatoa are chemolithotrophic sulfur-oxidizers, using reduced sulfur species as an energy source. They live in sulfur-rich environments such as soil, both marine and freshwater, in the deep sea hydrothermal vents, and in polluted marine environments. In association with other sulfur bacteria, e.g. Thiothrix, they can form biofilms that are visible to the naked eye as mats of long white filaments; the white color is due to sulfur globules stored inside the cells.

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.

<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. All members of this class are Gram-negative. It is the most phylogenetically and physiologically diverse class of the Pseudomonadota.

Biomining ( phytomining) is the concept of extracting metals from ores and other solid materials typically using prokaryotes, fungi or plants (phytoextraction. These organisms secrete organic compounds that chelate metals from the environment. The proposed technology is often aimed at extraction of iron, copper, zinc, gold, uranium, and thorium. Large chemostats of microbes can be grown to leach metals from their media. If it were practical, biomining would be an environmentally friendly alternative to traditional mining.

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.

<span class="mw-page-title-main">Bacterial anaerobic corrosion</span>

Bacterial anaerobic corrosion is the bacterially-induced oxidation of metals. Corrosion of metals typically alters the metal to a form that is more stable. Thus, bacterial anaerobic corrosion typically occurs in conditions favorable to the corrosion of the underlying substrate. In humid, anoxic conditions the corrosion of metals occurs as a result of a redox reaction. This redox reaction generates molecular hydrogen from local hydrogen ions. Conversely, anaerobic corrosion occurs spontaneously. Anaerobic corrosion primarily occurs on metallic substrates but may also occur on concrete.

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

The class Zetaproteobacteria is the sixth and most recently described class of the Pseudomonadota. 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, which is an iron-oxidizing neutrophilic chemolithoautotroph originally isolated from Kamaʻehuakanaloa Seamount in 1996 (post-eruption). 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.

<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.

Leptothrix cholodnii is a bacterium from the genus Leptothrix, which has the ability to oxidize Fe(II).. They were previously known as Leptothrix discophora SP-6. They are fast-growing metal oxidizers in iron-rich environments. These environments include freshwater bodies characterized with neutral to slightly acidic pH, oxygen gradients and organic matter. Examples of these sites include freshwater streams and wetlands, iron seeps, water pipes, surface of sediments Their growth under suitable conditions is easily recognized with fluffy microbial mats, surface biofilms made up of oxidized Fe and Mn minerals with orange to dark brown color. They can oxidize both Fe(II) and Mn(II). Leptothrix cholodnii SP-6 is a member of this group with an isolate and sheath-former under laboratory conditions.

Arsenate-reducing bacteria are bacteria which reduce arsenates. Arsenate-reducing bacteria are ubiquitous in arsenic-contaminated groundwater (aqueous environment). Arsenates are salts or esters of arsenic acid (H3AsO4), consisting of the ion AsO43−. They are moderate oxidizers that can be reduced to arsenites and to arsine. Arsenate can serve as a respiratory electron acceptor for oxidation of organic substrates and H2S or H2. Arsenates occur naturally in minerals such as adamite, alarsite, legrandite, and erythrite, and as hydrated or anhydrous arsenates. Arsenates are similar to phosphates since arsenic (As) and phosphorus (P) occur in group 15 (or VA) of the periodic table. Unlike phosphates, arsenates are not readily lost from minerals due to weathering. They are the predominant form of inorganic arsenic in aqueous aerobic environments. On the other hand, arsenite is more common in anaerobic environments, more mobile, and more toxic than arsenate. Arsenite is 25–60 times more toxic and more mobile than arsenate under most environmental conditions. Arsenate can lead to poisoning, since it can replace inorganic phosphate in the glyceraldehyde-3-phosphate --> 1,3-biphosphoglycerate step of glycolysis, producing 1-arseno-3-phosphoglycerate instead. Although glycolysis continues, 1 ATP molecule is lost. Thus, arsenate is toxic due to its ability to uncouple glycolysis. Arsenate can also inhibit pyruvate conversion into acetyl-CoA, thereby blocking the TCA cycle, resulting in additional loss of ATP.

Dissimilatory metal-reducing microorganisms are a group of microorganisms (both bacteria and archaea) that can perform anaerobic respiration utilizing a metal as terminal electron acceptor rather than molecular oxygen (O2), which is the terminal electron acceptor reduced to water (H2O) in aerobic respiration. The most common metals used for this end are iron [Fe(III)] and manganese [Mn(IV)], which are reduced to Fe(II) and Mn(II) respectively, and most microorganisms that reduce Fe(III) can reduce Mn(IV) as well. But other metals and metalloids are also used as terminal electron acceptors, such as vanadium [V(V)], chromium [Cr(VI)], molybdenum [Mo(VI)], cobalt [Co(III)], palladium [Pd(II)], gold [Au(III)], and mercury [Hg(II)].

<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.

References

  1. Andrews, Simon; Norton, Ian; Salunkhe, Arvindkumar S.; Goodluck, Helen; Aly, Wafaa S.M.; Mourad-Agha, Hanna; Cornelis, Pierre (2013). "Control of Iron Metabolism in Bacteria". In Banci (ed.). Metallomics and the Cell. Metal Ions in Life Sciences. Vol. 12. Springer. pp. 203–39. doi:10.1007/978-94-007-5561-1_7. ISBN   978-94-007-5560-4. PMID   23595674. electronic-book ISBN   978-94-007-5561-1 ISSN   1559-0836 electronic- ISSN   1868-0402
  2. Alth, Max; Alth, Charlotte (1984). Constructing and Maintaining Your Well & Septic System. Blue Ridge Summit, Pennsylvania: Tab Books. p. 20. ISBN   0-8306-0654-8.
  3. Krauskopf, Konrad B. "Introduction to Geochemistry" McGraw-Hill (1979) ISBN   0-07-035447-2 p.213
  4. Sawyer, Clair N., and McCarty, Perry L. "Chemistry for Sanitary Engineers" McGraw-Hill (1967) ISBN   0-07-054970-2 pp.446-447
  5. Krauskopf, Konrad B. "Introduction to Geochemistry" McGraw-Hill (1979) ISBN   0-07-035447-2 p.544
  6. Sawyer, Clair N., and McCarty, Perry L. "Chemistry for Sanitary Engineers" McGraw-Hill (1967) ISBN   0-07-054970-2 p.459
  7. 1 2 3 4 5 6 7 Madigan, Michael T.; Martinko, John M.; Stahl, David A.; Clark, David P. (2012). Brock biology of microorganisms (13th ed.). Boston: Benjamim Cummings. p. 1155. ISBN   978-0-321-64963-8.
  8. Bruslind, Linda (2019-08-01). "Chemolithotrophy & Nitrogen Metabolism".{{cite journal}}: Cite journal requires |journal= (help)
  9. 1 2 3 4 5 6 7 Hegler, Florian; Posth, Nicole R.; Jiang, Jie; Kappler, Andreas (1 November 2008). "Physiology of phototrophic iron(II)-oxidizing bacteria: implications for modern and ancient environments". FEMS Microbiology Ecology. 66 (2): 250–260. doi: 10.1111/j.1574-6941.2008.00592.x . ISSN   0168-6496. PMID   18811650.
  10. 1 2 3 4 5 6 Hedrich, S.; Schlomann, M.; Johnson, D. B. (21 April 2011). "The iron-oxidizing proteobacteria". Microbiology. 157 (6): 1551–1564. doi: 10.1099/mic.0.045344-0 . PMID   21511765.
  11. Caiazza, N. C.; Lies, D. P.; Newman, D. K. (10 August 2007). "Phototrophic Fe(II) Oxidation Promotes Organic Carbon Acquisition by Rhodobacter capsulatus SB1003". Applied and Environmental Microbiology. 73 (19): 6150–6158. Bibcode:2007ApEnM..73.6150C. doi:10.1128/AEM.02830-06. PMC   2074999 . PMID   17693559. S2CID   6110532.
  12. 1 2 Walter, Xavier A.; Picazo, Antonio; Miracle, Maria R.; Vicente, Eduardo; Camacho, Antonio; Aragno, Michel; Zopfi, Jakob (2014). "Phototrophic Fe(II)-oxidation in the chemocline of a ferruginous meromictic lake". Frontiers in Microbiology. 5: 9. doi: 10.3389/fmicb.2014.00713 . ISSN   1664-302X. PMC   4258642 . PMID   25538702.
  13. 1 2 3 4 Weber, Karrie A.; Achenbach, Laurie A.; Coates, John D. (October 2006). "Microorganisms pumping iron: anaerobic microbial iron oxidation and reduction" (PDF). Nature Reviews Microbiology. 4 (10): 752–764. doi:10.1038/nrmicro1490. ISSN   1740-1534. PMID   16980937. S2CID   91320892. Archived from the original (PDF) on 2019-12-03.
  14. Scholz, Florian; Löscher, Carolin R.; Fiskal, Annika; Sommer, Stefan; Hensen, Christian; Lomnitz, Ulrike; Wuttig, Kathrin; Göttlicher, Jörg; Kossel, Elke; Steininger, Ralph; Canfield, Donald E. (November 2016). "Nitrate-dependent iron oxidation limits iron transport in anoxic ocean regions". Earth and Planetary Science Letters. 454: 272–281. Bibcode:2016E&PSL.454..272S. doi:10.1016/j.epsl.2016.09.025.
  15. 1 2 Weber, Karrie A.; Pollock, Jarrod; Cole, Kimberly A.; O'Connor, Susan M.; Achenbach, Laurie A.; Coates, John D. (1 January 2006). "Anaerobic Nitrate-Dependent Iron(II) Bio-Oxidation by a Novel Lithoautotrophic Betaproteobacterium, Strain 2002". Applied and Environmental Microbiology. 72 (1): 686–694. Bibcode:2006ApEnM..72..686W. doi:10.1128/AEM.72.1.686-694.2006. ISSN   0099-2240. PMC   1352251 . PMID   16391108.
  16. 1 2 Muehe, EM; Gerhardt, S; Schink, B; Kappler, A (December 2009). "Ecophysiology and the energetic benefit of mixotrophic Fe(II) oxidation by various strains of nitrate-reducing bacteria". FEMS Microbiology Ecology. 70 (3): 335–43. doi: 10.1111/j.1574-6941.2009.00755.x . PMID   19732145.
  17. Scholz, Florian; Löscher, Carolin R.; Fiskal, Annika; Sommer, Stefan; Hensen, Christian; Lomnitz, Ulrike; Wuttig, Kathrin; Göttlicher, Jörg; Kossel, Elke; Steininger, Ralph; Canfield, Donald E. (2016). "Nitrate-dependent iron oxidation limits iron transport in anoxic ocean regions". Earth and Planetary Science Letters. 454: 272–281. Bibcode:2016E&PSL.454..272S. doi:10.1016/j.epsl.2016.09.025.
  18. 1 2 3 4 5 6 Emerson, David; Fleming, Emily J.; McBeth, Joyce M. (13 October 2010). "Iron-Oxidizing Bacteria: An Environmental and Genomic Perspective". Annual Review of Microbiology. 64 (1): 561–583. doi:10.1146/annurev.micro.112408.134208. PMID   20565252.
  19. Chan, Clara S; Fakra, Sirine C; Emerson, David; Fleming, Emily J; Edwards, Katrina J (2010-11-25). "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. ISSN   1751-7362. PMC   3105749 . PMID   21107443.
  20. 1 2 Emerson, David (2016). "The Irony of Iron–Biogenic Iron Oxides as an Iron Source to the Ocean". Frontiers in Microbiology. 6: 6. doi: 10.3389/fmicb.2015.01502 . PMC   4701967 . PMID   26779157.
  21. 1 2 McAllister, Sean M.; Moore, Ryan M.; Gartman, Amy; Luther, George W; Emerson, David; Chan, Clara S (30 January 2019). "The Fe(II)-Oxidizing Zetaproteobacteria: historical, ecological and genomic perspectives". FEMS Microbiology Ecology. 95 (4): 18. doi:10.1093/femsec/fiz015. PMC   6443915 . PMID   30715272.
  22. Henri, Pauline A; Rommevaux-Jestin, Céline; Lesongeur, Françoise; Mumford, Adam; Emerson, David; Godfroy, Anne; Ménez, Bénédicte (21 January 2016). "Structural Iron(II) of Basaltic Glass as an Energy Source for Zetaproteobacteria in an Abyssal Plain Environment, Off the Mid Atlantic Ridge". Frontiers in Microbiology. 6: 18. doi: 10.3389/fmicb.2015.01518 . PMC   4720738 . PMID   26834704.
  23. Makita, Hiroko (4 July 2018). "Iron-oxidizing bacteria in marine environments: recent progresses and future directions". World Journal of Microbiology and Biotechnology. 34 (8): 110. doi:10.1007/s11274-018-2491-y. ISSN   1573-0972. PMID   29974320. S2CID   49685224.
  24. Emerson, David; L. Moyer, Craig (June 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.
  25. Scott, Jarrod J.; Breier, John A.; Luther, George W.; Emerson, David; Duperron, Sebastien (11 March 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.
  26. Chan, CS; Fakra, SC; Emerson, D; Fleming, EJ; Edwards, KJ (April 2011). "Lithotrophic iron-oxidizing bacteria produce organic stalks to control mineral growth: implications for biosignature formation". The ISME Journal. 5 (4): 717–27. doi:10.1038/ismej.2010.173. PMC   3105749 . PMID   21107443.
  27. Hoegh-Guldberg, O.; Mumby, P. J.; Hooten, A. J.; Steneck, R. S.; Greenfield, P.; Gomez, E.; Harvell, C. D.; Sale, P. F.; Edwards, A. J.; Caldeira, K.; Knowlton, N. (2007-12-14). "Coral Reefs Under Rapid Climate Change and Ocean Acidification". Science. 318 (5857): 1737–1742. Bibcode:2007Sci...318.1737H. doi:10.1126/science.1152509. hdl: 1885/28834 . ISSN   0036-8075. PMID   18079392. S2CID   12607336.
  28. Deutsch, Curtis; Brix, Holger; Ito, Taka; Frenzel, Hartmut; Thompson, LuAnne (2011-06-09). "Climate-Forced Variability of Ocean Hypoxia". Science. 333 (6040): 336–339. Bibcode:2011Sci...333..336D. doi:10.1126/science.1202422. ISSN   0036-8075. PMID   21659566. S2CID   11752699.
  29. Emerson, David (2016-01-06). "The Irony of Iron – Biogenic Iron Oxides as an Iron Source to the Ocean". Frontiers in Microbiology. 6: 1502. doi: 10.3389/fmicb.2015.01502 . ISSN   1664-302X. PMC   4701967 . PMID   26779157.
  30. Hazan, Zadik; Zumeris, Jona; Jacob, Harold; Raskin, Hanan; Kratysh, Gera; Vishnia, Moshe; Dror, Naama; Barliya, Tilda; Mandel, Mathilda; Lavie, Gad (2006-12-01). "Strategies to prevent, curb and eliminate biofilm formation based on the characteristics of various periods in one biofilm life cycle". Antimicrobial Agents and Chemotherapy. 50 (12): 4144–4152. doi:10.1128/AAC.00418-06. PMC   1693972 . PMID   16940055.
  31. Ramasamy, Mohankandhasamy; Lee, Jintae (2016). "Recent Nanotechnology Approaches for Prevention and Treatment of Biofilm-Associated Infections on Medical Devices". BioMed Research International. 2016: 1851242. doi: 10.1155/2016/1851242 . ISSN   2314-6141. PMC   5107826 . PMID   27872845.
  32. Ma, Ruixiang; Hu, Xianli; Zhang, Xianzuo; Wang, Wenzhi; Sun, Jiaxuan; Su, Zheng; Zhu, Chen (2022). "Strategies to prevent, curb and eliminate biofilm formation based on the characteristics of various periods in one biofilm life cycle". Frontiers in Cellular and Infection Microbiology. 12: 1003033. doi: 10.3389/fcimb.2022.1003033 . ISSN   2235-2988. PMC   9534288 . PMID   36211965.
  33. "Information for you about Iron Bacteria & Well Water" (PDF). Information for You. 1 (1): 3. 2017.
  34. Barry, Dana M.; Kanematsu, Hideyuki (2015), Kanematsu, Hideyuki; Barry, Dana M. (eds.), "Physical Removal of Biofilm", Biofilm and Materials Science, Cham: Springer International Publishing, pp. 163–167, doi:10.1007/978-3-319-14565-5_20, ISBN   978-3-319-14565-5 , retrieved 2022-12-22