Redox gradient

Last updated
Depiction of common redox reactions in the environment. Adapted from figures by Zhang and Gorny. Redox pairs are listed with the oxidizer (electron acceptor) in red and the reducer (electron donator) in black. Redox Tower (2).png
Depiction of common redox reactions in the environment. Adapted from figures by Zhang and Gorny. Redox pairs are listed with the oxidizer (electron acceptor) in red and the reducer (electron donator) in black.
Relative favorability of redox reactions in marine sediments based on energy. Start points of arrows indicate energy associated with half-cell reaction. Lengths of arrows indicate an estimate of Gibb's free energy (DG) for the reaction where a higher DG is more energetically favorable (Adapted from Libes, 2011). 2Reduction reaction energetics.png
Relative favorability of redox reactions in marine sediments based on energy. Start points of arrows indicate energy associated with half-cell reaction. Lengths of arrows indicate an estimate of Gibb's free energy (ΔG) for the reaction where a higher ΔG is more energetically favorable (Adapted from Libes, 2011).

A redox gradient is a series of reduction-oxidation (redox) reactions sorted according to redox potential. [4] [5] The redox ladder displays the order in which redox reactions occur based on the free energy gained from redox pairs. [4] [5] [6] These redox gradients form both spatially and temporally as a result of differences in microbial processes, chemical composition of the environment, and oxidative potential. [5] [4] Common environments where redox gradients exist are coastal marshes, lakes, contaminant plumes, and soils. [1] [4] [5] [6]

Contents

The Earth has a global redox gradient with an oxidizing environment at the surface and increasingly reducing conditions below the surface. [4] Redox gradients are generally understood at the macro level, but characterization of redox reactions in heterogeneous environments at the micro-scale require further research and more sophisticated measurement techniques. [5] [1] [7] [6]

Measuring redox conditions

Redox conditions are measured according to the redox potential (Eh) in volts, which represents the tendency for electrons to transfer from an electron donor to an electron acceptor. Eh can be calculated using half reactions and the Nernst equation. [1] An Eh of zero represents the redox couple of the standard hydrogen electrode H+/H2, [8] a positive Eh indicates an oxidizing environment (electrons will be accepted), and a negative Eh indicates a reducing environment (electrons will be donated). [1] In a redox gradient, the most energetically favorable chemical reaction occurs at the “top” of the redox ladder and the least energetically favorable reaction occurs at the “bottom” of the ladder. [1]

Eh can be measured by collecting samples in the field and performing analyses in the lab, or by inserting an electrode into the environment to collect in situ measurements. [6] [5] [1] Typical environments to measure redox potential are in bodies of water, soils, and sediments, all of which can exhibit high levels of heterogeneity. [5] [1] Collecting a high number of samples can produce high spatial resolution, but at the cost of low temporal resolution since samples only reflect a singular a snapshot in time. [8] [1] [5] In situ monitoring can provide high temporal resolution by collecting continuous real-time measurements, but low spatial resolution since the electrode is in a fixed location. [1] [5]

Redox properties can also be tracked with high spatial and temporal resolution through the use of induced-polarization imaging, however, further research is needed to fully understand contributions of redox species to polarization. [6]

Environmental conditions

Redox gradients are commonly found in the environment as functions of both space and time, [9] [8] particularly in soils and aquatic environments. [8] [6] Gradients are caused by varying physiochemical properties including availability of oxygen, soil hydrology, chemical species present, and microbial processes. [1] [4] [9] [8] Specific environments that are commonly characterized by redox gradients include waterlogged soils, wetlands, [8] contaminant plumes, [9] [4] and marine pelagic and hemipelagic sediments. [4]

The following is a list of common reactions that occur in the environment in order from oxidizing to reducing (organisms performing the reaction in parentheses): [1]

  1. Aerobic respiration (aerobes: aerobic organisms)
  2. Denitrification (denitrifiers: denitrifying bacteria)
  3. Manganese reduction (Manganese reducers)
  4. Iron reduction (iron reducers: iron-reducing bacteria)
  5. Sulfate reduction (sulfate reducers: Sulfur-reducing bacteria)
  6. Methanogenesis (methanogens)

Aquatic environments

Redox gradients form in water columns and their sediments. Varying levels of oxygen (oxic, suboxic, hypoxic) within the water column alter redox chemistry and which redox reactions can occur. [10] Development of oxygen minimum zones also contributes to formation of redox gradients.

Benthic sediments exhibit redox gradients produced by variations in mineral composition, organic matter availability, structure, and sorption dynamics. [5] Limited transport of dissolved electrons through subsurface sediments, combined with varying pore sizes of sediments creates significant heterogeneity in benthic sediments. [5] Oxygen availability in sediments determines which microbial respiration pathways can occur, resulting in a vertical stratification of redox processes as oxygen availability decreases with depth. [5]

Terrestrial environments

Soil Eh is also largely a function of hydrological conditions. [1] [8] [6] In the event of a flood, saturated soils can shift from oxic to anoxic, creating a reducing environment as anaerobic microbial processes dominate. [1] [8] Moreover, small anoxic hotspots may develop within soil pore spaces, creating reducing conditions. [6] With time, the starting Eh of a soil can be restored as water drains and the soil dries out. [1] [8] Soils with redox gradients formed by ascending groundwater are classified as gleysols, while soils with gradients formed by stagnant water are classified as stagnosols and planosols.

Soil Eh generally ranges from −300 to +900 mV. [8] The table below summarizes typical Eh values for various soil conditions: [1] [8]

Soil conditionsTypical Eh range (mV) [1] [8]
WaterloggedEh < +250
Aerated – moderately reduced+100 < Eh < +400
Aerated – reduced−100 < Eh < +100
Aerated – highly reduced−300 < Eh < −100
Cultivated+300 < Eh < +500

Generally accepted Eh limits that are tolerable by plants are +300 mV < Eh < +700 mV. [8] 300 mV is the boundary value that separates aerobic from anaerobic conditions in wetland soils. [1] Redox potential (Eh) is also closely tied to pH, and both have significant influence on the function of soil-plant-microorganism systems. [1] [8] The main source of electrons in soil is organic matter. [8] Organic matter consumes oxygen as it decomposes, resulting in reducing soil conditions and lower Eh. [8]

Role of microorganisms

Redox gradients form based on resource availability and physiochemical conditions (pH, salinity, temperature) and support stratified communities of microbes. [1] [5] [9] [8] [7] Microbes carry out differing respiration processes (methanogenesis, sulfate reduction, etc.) based on the conditions around them and further amplify redox gradients present in the environment. [9] [1] [8] However, distribution of microorganisms cannot solely be determined from thermodynamics (redox ladder), but is also influenced by ecological and physiological factors. [6] [5]

Redox gradients form along contaminant plumes, in both aquatic and terrestrial settings, as a function of the contaminant concentration and the impacts it has on relevant chemical processes and microbial communities. [1] [9] The highest rates of organic pollutant degradation along a redox gradient are found at the oxic-anoxic interface. [1] In groundwater, this oxic-anoxic environment is referred to as the capillary fringe, where the water table meets soil and fills empty pores. Because this transition zone is both oxic and anoxic, electron acceptors and donors are in high abundance and there is a high level of microbial activity, leading to the highest rates of contaminant biodegradation. [1] [9]

Benthic sediments are heterogeneous in nature and subsequently exhibit redox gradients. [5] Due to this heterogeneity, gradients of reducing and oxidizing chemical species do not always overlap enough to support electron transport needs of niche microbial communities. [5] Cable bacteria have been characterized as sulfide-oxidizing bacteria that assist in connecting these areas of undersupplied and excess electrons to complete the electron transport for otherwise unavailable redox reactions. [5]

Biofilms, found in tidal flats, glaciers, hydrothermal vents, and at the bottoms of aquatic environments, also exhibit redox gradients. [5] The community of microbes—often metal- or sulfate-reducing bacteria—produces redox gradients on the micrometer scale as a function of spatial physiochemical variability. [5]

See sulfate-methane transition zone for coverage of microbial processes in SMTZs.

See also

Related Research Articles

<span class="mw-page-title-main">Redox</span> Chemical reaction in which oxidation states of atoms are changed

Redox is a type of chemical reaction in which the oxidation states of a reactant change. Oxidation is the loss of electrons or an increase in the oxidation state, while reduction is the gain of electrons or a decrease in the oxidation state.

An electron transport chain (ETC) is a series of protein complexes and other molecules that transfer electrons from electron donors to electron acceptors via redox reactions (both reduction and oxidation occurring simultaneously) and couples this electron transfer with the transfer of protons (H+ ions) across a membrane. Many of the enzymes in the electron transport chain are embedded within the membrane.

<span class="mw-page-title-main">Cellular respiration</span> Process to convert glucose to ATP in cells

Cellular respiration is the process by which biological fuels are oxidized in the presence of an inorganic electron acceptor, such as oxygen, to drive the bulk production of adenosine triphosphate (ATP), which contains energy. Cellular respiration may be described as a set of metabolic reactions and processes that take place in the cells of organisms to convert chemical energy from nutrients into ATP, and then release waste products.

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.

The pedosphere is the outermost layer of the Earth that is composed of soil and subject to soil formation processes. It exists at the interface of the lithosphere, atmosphere, hydrosphere and biosphere. The pedosphere is the skin of the Earth and only develops when there is a dynamic interaction between the atmosphere, biosphere, lithosphere and the hydrosphere. The pedosphere is the foundation of terrestrial life on Earth.

<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.

<span class="mw-page-title-main">Sediment–water interface</span> The boundary between bed sediment and the overlying water column

In oceanography and limnology, the sediment–water interface is the boundary between bed sediment and the overlying water column. The term usually refers to a thin layer of water at the very surface of sediments on the seafloor. In the ocean, estuaries, and lakes, this layer interacts with the water above it through physical flow and chemical reactions mediated by the micro-organisms, animals, and plants living at the bottom of the water body. The topography of this interface is often dynamic, as it is affected by physical processes and biological processes. Physical, biological, and chemical processes occur at the sediment-water interface as a result of a number of gradients such as chemical potential gradients, pore water gradients, and oxygen gradients.

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

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

In biogeochemistry, remineralisation refers to the breakdown or transformation of organic matter into its simplest inorganic forms. These transformations form a crucial link within ecosystems as they are responsible for liberating the energy stored in organic molecules and recycling matter within the system to be reused as nutrients by other organisms.

<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.

Anoxic waters are areas of sea water, fresh water, or groundwater that are depleted of dissolved oxygen. The US Geological Survey defines anoxic groundwater as those with dissolved oxygen concentration of less than 0.5 milligrams per litre. Anoxic waters can be contrasted with hypoxic waters, which are low in dissolved oxygen. This condition is generally found in areas that have restricted water exchange.

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.

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.

Dissimilatory nitrate reduction to ammonium (DNRA), also known as nitrate/nitrite ammonification, is the result of anaerobic respiration by chemoorganoheterotrophic microbes using nitrate (NO3) as an electron acceptor for respiration. In anaerobic conditions microbes which undertake DNRA oxidise organic matter and use nitrate (rather than oxygen) as an electron acceptor, reducing it to nitrite, and then to ammonium (NO3 → NO2 → NH4+).

Euxinia or euxinic conditions occur when water is both anoxic and sulfidic. This means that there is no oxygen (O2) and a raised level of free hydrogen sulfide (H2S). Euxinic bodies of water are frequently strongly stratified; have an oxic, highly productive, thin surface layer; and have anoxic, sulfidic bottom water. The word "euxinia" is derived from the Greek name for the Black Sea (Εὔξεινος Πόντος (Euxeinos Pontos)) which translates to "hospitable sea". Euxinic deep water is a key component of the Canfield ocean, a model of oceans during part of the Proterozoic eon (a part specifically known as the Boring Billion) proposed by Donald Canfield, an American geologist, in 1998. There is still debate within the scientific community on both the duration and frequency of euxinic conditions in the ancient oceans. Euxinia is relatively rare in modern bodies of water, but does still happen in places like the Black Sea and certain fjords.

<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).

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.

An oxygen minimum zone (OMZ) is characterized as an oxygen-deficient layer in the world's oceans. Typically found between 200m to 1500m deep below regions of high productivity, such as the western coasts of continents. OMZs can be seasonal following the spring-summer upwelling season. Upwelling of nutrient-rich water leads to high productivity and labile organic matter, that is respired by heterotrophs as it sinks down the water column. High respiration rates deplete the oxygen in the water column to concentrations of 2 mg/L or less forming the OMZ. OMZs are expanding, with increasing ocean deoxygenation. Under these oxygen-starved conditions, energy is diverted from higher trophic levels to microbial communities that have evolved to use other biogeochemical species instead of oxygen, these species include Nitrate, Nitrite, Sulphate etc. Several Bacteria and Archea have adapted to live in these environments by using these alternate chemical species and thrive. The most abundant phyla in OMZs are Pseudomonadota, Bacteroidota, Actinomycetota, and Planctomycetota.

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 Zhang, Zengyu; Furman, Alex (2021). "Soil redox dynamics under dynamic hydrologic regimes - A review". Science of the Total Environment. 763: 143026. Bibcode:2021ScTEn.763n3026Z. doi:10.1016/j.scitotenv.2020.143026. ISSN   0048-9697. PMID   33143917. S2CID   226249448.
  2. Gorny, J.; Billon, G.; Lesven, L.; Dumoulin, D.; Madé, B.; Noiriel, C. (2015). "Arsenic behavior in river sediments under redox gradient: a review". The Science of the Total Environment. 505: 423–434. doi:10.1016/j.scitotenv.2014.10.011. PMID   25461044. S2CID   24877798.
  3. Libes, Susan (2009). Introduction to marine biogeochemistry. Amsterdam Boston: Elsevier/Academic Press. ISBN   978-0-08-091664-4. OCLC   643573176.
  4. 1 2 3 4 5 6 7 8 Borch, Thomas; Kretzschmar, Ruben; Kappler, Andreas; Cappellen, Philippe Van; Ginder-Vogel, Matthew; Voegelin, Andreas; Campbell, Kate (2009). "Biogeochemical Redox Processes and their Impact on Contaminant Dynamics". Environmental Science & Technology. 44 (1). American Chemical Society (ACS): 15–23. doi:10.1021/es9026248. ISSN   0013-936X. PMID   20000681. S2CID   206997593.
  5. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 Lau, Maximilian Peter; Niederdorfer, Robert; Sepulveda-Jauregui, Armando; Hupfer, Michael (2018). "Synthesizing redox biogeochemistry at aquatic interfaces". Limnologica. 68: 59–70. doi: 10.1016/j.limno.2017.08.001 .
  6. 1 2 3 4 5 6 7 8 9 Peiffer, S.; Kappler, A.; Haderlein, S. B.; Schmidt, C.; Byrne, J. M.; Kleindienst, S.; Vogt, C.; Richnow, H. H.; Obst, M.; Angenent, L. T.; Bryce, C. (2021). "A biogeochemical–hydrological framework for the role of redox-active compounds in aquatic systems". Nature Geoscience. 14 (5): 264–272. Bibcode:2021NatGe..14..264P. doi:10.1038/s41561-021-00742-z. ISSN   1752-0894. S2CID   233876038.
  7. 1 2 Zakem, Emily J.; Polz, Martin F.; Follows, Michael J. (2020). "Redox-informed models of global biogeochemical cycles". Nature Communications. 11 (1): 5680. Bibcode:2020NatCo..11.5680Z. doi:10.1038/s41467-020-19454-w. ISSN   2041-1723. PMC   7656242 . PMID   33173062.
  8. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Husson, Olivier (2013). "Redox potential (Eh) and pH as drivers of soil/plant/microorganism systems: a transdisciplinary overview pointing to integrative opportunities for agronomy". Plant and Soil. 362 (1–2): 389–417. doi: 10.1007/s11104-012-1429-7 . ISSN   0032-079X. S2CID   17059599.
  9. 1 2 3 4 5 6 7 Vodyanitskii, Yu N. (2016). "Biochemical processes in soil and groundwater contaminated by leachates from municipal landfills (Mini review)". Annals of Agrarian Science. 14 (3): 249–256. doi: 10.1016/j.aasci.2016.07.009 . ISSN   1512-1887.
  10. Rue, Eden L.; Smith, Geoffrey J.; Cutter, Gregory A.; Bruland, Kenneth W. (1997). "The response of trace element redox couples to suboxic conditions in the water column". Deep Sea Research Part I: Oceanographic Research Papers. 44 (1): 113–134. Bibcode:1997DSRI...44..113R. doi:10.1016/S0967-0637(96)00088-X.