Mantle oxidation state

Last updated
Oxygen fugacity range where common cation pairs dominate.Data for plotting are from Shearer et al., (2006). IW represents Iron-Wustite buffer and QFM represents Quartz-Fayalite-Magnetite buffer. Oxygen fugacity range where cation pairs dominate.png
Oxygen fugacity range where common cation pairs dominate.Data for plotting are from Shearer et al., (2006). IW represents Iron-Wüstite buffer and QFM represents Quartz-Fayalite-Magnetite buffer.

Mantle oxidation state (redox state) applies the concept of oxidation state in chemistry to the study of the Earth's mantle. The chemical concept of oxidation state mainly refers to the valence state of one element, while mantle oxidation state provides the degree of decreasing of increasing valence states of all polyvalent elements in mantle materials confined in a closed system. The mantle oxidation state is controlled by oxygen fugacity and can be benchmarked by specific groups of redox buffers.

Contents

Mantle oxidation state changes because of the existence of polyvalent elements (elements with more than one valence state, e.g. Fe, Cr, V, Ti, Ce, Eu, C and others). Among them, Fe is the most abundant (≈8 wt% of the mantle [2] ) and its oxidation state largely reflects the oxidation state of mantle. Examining the valence state of other polyvalent elements could also provide the information of mantle oxidation state.

It is well known that the oxidation state can influence the partitioning behavior of elements [3] [4] and liquid water [5] between melts and minerals, the speciation of C-O-H-bearing fluids and melts, [6] as well as transport properties like electrical conductivity and creep. [5]

The formation of diamond requires both reaching high pressures and high temperatures and a carbon source. The most common carbon source in deep Earth is not elemental carbon and redox reactions need to be involved in diamond formation. Examining the oxidation state can help us predict the P-T conditions of diamond formation and elucidate the origin of deep diamonds. [7]

Thermodynamic description of oxidation state

Mantle oxidation state can be quantified as the oxygen fugacity () of the system within the framework of thermodynamics. A higher oxygen fugacity implies a more oxygen-rich and more oxidized environment. At each given pressure-temperature conditions, for any compound or element M that bears the potential to be oxidized by oxygen [8] [9]

For example, if M is Fe, the redox equilibrium reaction can be Fe+1/2O2=FeO; if M is FeO, the redox equilibrium reaction can be 2FeO+1/2O2=Fe2O3.

Gibbs energy change associated with this reaction is therefore

Along each isotherm, the partial derivation of ΔG with respect to P is ΔV,

.[ citation needed ]

Combining the 2 equations above,

.

Therefore,

(note that ln(e as the base) changed to log(10 as the base) in this formula.

For a closed system, there might exist more than one of these equilibrium oxidation reactions, but since all these reactions share a same , examining one of them would allow extraction of oxidation state of the system.

Pressure effect on oxygen fugacity

The physics and chemistry of mantle largely depend on pressure. As mantle minerals are compressed, they are transformed into other minerals at certain depths. Seismic observations of velocity discontinuities and experimental simulations on phase boundaries both verified the structure transformations within the mantle. As such, the mantle can be further divided into three layers with distinct mineral compositions.

Mantle Mineral Composition [10]
Mantle LayerDepthPressureMajor Minerals
Upper Mantle≈10–410 km≈1-13 GPa Olivine, Orthopyroxene, Clinopyroxene, Garnet
Transition Zone410–660 km13-23 GPa Wadsleyite, Ringwoodite, Majoritic Garnet
Lower Mantle660–2891 km23-129 GPa Ferropericlase, Bridgmanite, Ca-perovskite

Since mantle mineral composition changes, the mineral hosting environment for polyvalent elements also alters. For each layer, the mineral combination governing the redox reactions is unique and will be discussed in detailed below.

Upper mantle

Between depths of 30 and 60 km, oxygen fugacity is mainly controlled by Olivine-Orthopyroxene-Spinel oxidation reaction.

Under deeper upper mantle conditions, Olivine-Orthopyroxene-Garnet oxygen barometer [11] is the redox reaction that is used to calibrate oxygen fugacity.

In this reaction, 4 mole of ferrous ions were oxidized to ferric ions and the other 2 mole of ferrous ions remain unchanged.

Transition zone

Garnet-Garnet [12] reaction can be used to estimate the redox state of transition zone.

Garnet, a major mineral in the transition zone, controls the oxidation state there. Garnet-Group-215473.jpg
Garnet, a major mineral in the transition zone, controls the oxidation state there.

A recent study [12] showed that the oxygen fugacity of transition referred from Garnet-Garnet reaction is -0.26 to +3 relative to the Fe-FeO (IW, iron- wütstite) oxygen buffer.

Lower mantle

Disproportionation of ferrous iron at lower mantle conditions also affect the mantle oxidation state. This reaction is different from the reactions mentioned above as it does not incorporate the participation of free oxygen.

, [5] [13]

FeO resides in the form of ferropericlase (Fp) and Fe2O3 resides in the form of bridgmanite (Bdg). There is no oxygen fugacity change associated with the reaction. However, as the reaction products differ in density significantly, the metallic iron phase could descend downwards to the Earth's core and get separated from the mantle. In this case, the mantle loses metallic iron and becomes more oxidized.

Implications for diamond formation

Diamond formed in the Earth's interior Diamond-diamond macle1.jpg
Diamond formed in the Earth's interior

The equilibrium reaction involving diamond is

Carbon Cycle involving deep Earth Flux of crustal material in the mantle.jpg
Carbon Cycle involving deep Earth

.

Examining the oxygen fugacity of the upper mantle and transition enables us to compare it with the conditions (equilibrium reaction shown above) required for diamond formation. The results show that the is usually 2 units lower than the carbonate-carbon reaction [12] which means favoring the formation of diamond at transition zone conditions.

It has also been reported that pH decrease would also facilitate the formation of diamond in Mantle conditions. [14]

where the subscript aq means 'aqueous', implying H2 is dissolved in the solution.

Deep diamonds have become important windows to look into the mineralogy of the Earth's interior. Minerals not stable at the surface could possibly be found within inclusions of superdeep diamonds [15] —implying they were stable where these diamond crystallized. Because of the hardness of diamonds, the high pressure environment is retained even after transporting to the surface. So far, these superdeep minerals brought by diamonds include ringwoodite, [16] ice-VII, [17] cubic δ-N2 [18] and Ca-perovskite. [19]

See also

Related Research Articles

<span class="mw-page-title-main">Stoichiometry</span> Calculation of relative weights of reactants and products in chemical reactions

Stoichiometry is the relationship between the weights of reactants and products before, during, and following chemical reactions.

In chemistry, a half reaction is either the oxidation or reduction reaction component of a redox reaction. A half reaction is obtained by considering the change in oxidation states of individual substances involved in the redox reaction. Often, the concept of half reactions is used to describe what occurs in an electrochemical cell, such as a Galvanic cell battery. Half reactions can be written to describe both the metal undergoing oxidation and the metal undergoing reduction.

In electrochemistry, the Nernst equation is a chemical thermodynamical relationship that permits the calculation of the reduction potential of a reaction from the standard electrode potential, absolute temperature, the number of electrons involved in the redox reaction, and activities of the chemical species undergoing reduction and oxidation respectively. It was named after Walther Nernst, a German physical chemist who formulated the equation.

In chemistry, a reducing agent is a chemical species that "donates" an electron to an electron recipient.

<span class="mw-page-title-main">Wüstite</span> Iron(II) oxide (FeO) mineral formed under reducing conditions

Wüstite is a mineral form of mostly iron(II) oxide found with meteorites and native iron. It has a grey colour with a greenish tint in reflected light. Wüstite crystallizes in the isometric-hexoctahedral crystal system in opaque to translucent metallic grains. It has a Mohs hardness of 5 to 5.5 and a specific gravity of 5.88. Wüstite is a typical example of a non-stoichiometric compound.

In electrochemistry, the standard hydrogen electrode, is a redox electrode which forms the basis of the thermodynamic scale of oxidation-reduction potentials. Its absolute electrode potential is estimated to be 4.44 ± 0.02 V at 25 °C, but to form a basis for comparison with all other electrochemical reactions, hydrogen's standard electrode potential is declared to be zero volts at any temperature. Potentials of all other electrodes are compared with that of the standard hydrogen electrode at the same temperature.

<span class="mw-page-title-main">Oxygen cycle</span> Biogeochemical cycle of oxygen

Oxygen cycle refers to the movement of oxygen through the atmosphere (air), biosphere (plants and animals) and the lithosphere (the Earth’s crust). The oxygen cycle demonstrates how free oxygen is made available in each of these regions, as well as how it is used. The oxygen cycle is the biogeochemical cycle of oxygen atoms between different oxidation states in ions, oxides, and molecules through redox reactions within and between the spheres/reservoirs of the planet Earth. The word oxygen in the literature typically refers to the most common oxygen allotrope, elemental/diatomic oxygen (O2), as it is a common product or reactant of many biogeochemical redox reactions within the cycle. Processes within the oxygen cycle are considered to be biological or geological and are evaluated as either a source (O2 production) or sink (O2 consumption).

The abiogenic petroleum origin hypothesis proposes that most of earth's petroleum and natural gas deposits were formed inorganically, commonly known as abiotic oil. Scientific evidence overwhelmingly supports a biogenic origin for most of the world's petroleum deposits. Mainstream theories about the formation of hydrocarbons on earth point to an origin from the decomposition of long-dead organisms, though the existence of hydrocarbons on extraterrestrial bodies like Saturn's moon Titan indicates that hydrocarbons are sometimes naturally produced by inorganic means. A historical overview of theories of the abiogenic origins of hydrocarbons has been published.

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.

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.

Redox potential is a measure of the tendency of a chemical species to acquire electrons from or lose electrons to an electrode and thereby be reduced or oxidised respectively. Redox potential is expressed in volts (V). Each species has its own intrinsic redox potential; for example, the more positive the reduction potential, the greater the species' affinity for electrons and tendency to be reduced.

<span class="mw-page-title-main">Pourbaix diagram</span> Plot of thermodynamically stable phases of an aqueous electrochemical system

In electrochemistry, and more generally in solution chemistry, a Pourbaix diagram, also known as a potential/pH diagram, EH–pH diagram or a pE/pH diagram, is a plot of possible thermodynamically stable phases of an aqueous electrochemical system. Boundaries (50 %/50 %) between the predominant chemical species are represented by lines. As such a Pourbaix diagram can be read much like a standard phase diagram with a different set of axes. Similarly to phase diagrams, they do not allow for reaction rate or kinetic effects. Beside potential and pH, the equilibrium concentrations are also dependent upon, e.g., temperature, pressure, and concentration. Pourbaix diagrams are commonly given at room temperature, atmospheric pressure, and molar concentrations of 10−6 and changing any of these parameters will yield a different diagram.

In theoretical chemistry, Marcus theory is a theory originally developed by Rudolph A. Marcus, starting in 1956, to explain the rates of electron transfer reactions – the rate at which an electron can move or jump from one chemical species (called the electron donor) to another (called the electron acceptor). It was originally formulated to address outer sphere electron transfer reactions, in which the two chemical species only change in their charge with an electron jumping (e.g. the oxidation of an ion like Fe2+/Fe3+), but do not undergo large structural changes. It was extended to include inner sphere electron transfer contributions, in which a change of distances or geometry in the solvation or coordination shells of the two chemical species is taken into account (the Fe-O distances in Fe(H2O)2+ and Fe(H2O)3+ are different).

<span class="mw-page-title-main">Mineral redox buffer</span> Geochemical assemblage

In geology, a redox buffer is an assemblage of minerals or compounds that constrains oxygen fugacity as a function of temperature. Knowledge of the redox conditions (or equivalently, oxygen fugacities) at which a rock forms and evolves can be important for interpreting the rock history. Iron, sulfur, and manganese are three of the relatively abundant elements in the Earth's crust that occur in more than one oxidation state. For instance, iron, the fourth most abundant element in the crust, exists as native iron, ferrous iron (Fe2+), and ferric iron (Fe3+). The redox state of a rock affects the relative proportions of the oxidation states of these elements and hence may determine both the minerals present and their compositions. If a rock contains pure minerals that constitute a redox buffer, then the oxygen fugacity of equilibration is defined by one of the curves in the accompanying fugacity-temperature diagram.

<span class="mw-page-title-main">Ferredoxin hydrogenase</span> Class of enzymes

In enzymology, ferredoxin hydrogenase, also referred to as [Fe-Fe]hydrogenase, H2 oxidizing hydrogenase, H2 producing hydrogenase, bidirectional hydrogenase, hydrogenase (ferredoxin), hydrogenlyase, and uptake hydrogenase, is found in Clostridium pasteurianum, Clostridium acetobutylicum,Chlamydomonas reinhardtii, and other organisms. The systematic name of this enzyme is hydrogen:ferredoxin oxidoreductase

<span class="mw-page-title-main">Solid oxide electrolyzer cell</span> Type of fuel cell

A solid oxide electrolyzer cell (SOEC) is a solid oxide fuel cell that runs in regenerative mode to achieve the electrolysis of water by using a solid oxide, or ceramic, electrolyte to produce hydrogen gas and oxygen. The production of pure hydrogen is compelling because it is a clean fuel that can be stored, making it a potential alternative to batteries, methane, and other energy sources. Electrolysis is currently the most promising method of hydrogen production from water due to high efficiency of conversion and relatively low required energy input when compared to thermochemical and photocatalytic methods.

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

For chemical reactions, the iron oxide cycle (Fe3O4/FeO) is the original two-step thermochemical cycle proposed for use for hydrogen production. It is based on the reduction and subsequent oxidation of iron ions, particularly the reduction and oxidation between Fe3+ and Fe2+. The ferrites, or iron oxide, begins in the form of a spinel and depending on the reaction conditions, dopant metals and support material forms either Wüstites or different spinels.

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

Electro-oxidation(EO or EOx), also known as anodic oxidation or electrochemical oxidation (EC), is a technique used for wastewater treatment, mainly for industrial effluents, and is a type of advanced oxidation process (AOP). The most general layout comprises two electrodes, operating as anode and cathode, connected to a power source. When an energy input and sufficient supporting electrolyte are provided to the system, strong oxidizing species are formed, which interact with the contaminants and degrade them. The refractory compounds are thus converted into reaction intermediates and, ultimately, into water and CO2 by complete mineralization.

In chemistry, the oxygen reduction reaction refers to the reduction half reaction whereby O2 is reduced to water or hydrogen peroxide. In fuel cells, the reduction to water is preferred because the current is higher. The oxygen reduction reaction is well demonstrated and highly efficient in nature.

References

  1. Shearer, C.K.; Papike, J.J.; Karner, J.M. (2006-10-01). "Pyroxene europium valence oxybarometer: Effects of pyroxene composition, melt composition, and crystallization kinetics". American Mineralogist. 91 (10): 1565–1573. Bibcode:2006AmMin..91.1565S. doi:10.2138/am.2006.2098. ISSN   0003-004X. S2CID   2080884.
  2. McDonough, W. F.; Sun, S. -s. (1995-03-01). "The composition of the Earth". Chemical Geology. Chemical Evolution of the Mantle. 120 (3): 223–253. Bibcode:1995ChGeo.120..223M. doi:10.1016/0009-2541(94)00140-4. ISSN   0009-2541.
  3. Fischer, Rebecca A.; Nakajima, Yoichi; Campbell, Andrew J.; Frost, Daniel J.; Harries, Dennis; Langenhorst, Falko; Miyajima, Nobuyoshi; Pollok, Kilian; Rubie, David C. (2015-10-15). "High pressure metal–silicate partitioning of Ni, Co, V, Cr, Si, and O". Geochimica et Cosmochimica Acta. 167: 177–194. Bibcode:2015GeCoA.167..177F. doi: 10.1016/j.gca.2015.06.026 . ISSN   0016-7037.
  4. Corgne, Alexandre; Keshav, Shantanu; Wood, Bernard J.; McDonough, William F.; Fei, Yingwei (2008). "Metal–silicate partitioning and constraints on core composition and oxygen fugacity during Earth accretion". Geochimica et Cosmochimica Acta. 72 (2): 574–589. Bibcode:2008GeCoA..72..574C. doi:10.1016/j.gca.2007.10.006.
  5. 1 2 3 Frost, Daniel J.; McCammon, Catherine A. (2008-04-29). "The Redox State of Earth's Mantle". Annual Review of Earth and Planetary Sciences. 36 (1): 389–420. Bibcode:2008AREPS..36..389F. doi:10.1146/annurev.earth.36.031207.124322. ISSN   0084-6597.
  6. Holloway, John R.; Blank, Jennifer G. (1994-12-31), "Chapter 6. APPLICATION OF EXPERIMENTAL RESULTS TO C-O-H SPECIES IN NATURAL MELTS", Volatiles in Magmas, De Gruyter, pp. 187–230, doi:10.1515/9781501509674-012, ISBN   9781501509674
  7. Luth, R. W.; Stachel, T. (2015). "Diamond formation — Where, when and how?". Lithos. Complete (220–223): 200–220. Bibcode:2015Litho.220..200S. doi:10.1016/j.lithos.2015.01.028. ISSN   0024-4937.
  8. Zhang, H.L.; Hirschmann, M.M.; Cottrell, E.; Withers, A.C. (2017). "Effect of pressure on Fe 3+ /ΣFe ratio in a mafic magma and consequences for magma ocean redox gradients". Geochimica et Cosmochimica Acta. 204: 83–103. Bibcode:2017GeCoA.204...83Z. doi: 10.1016/j.gca.2017.01.023 . ISSN   0016-7037.
  9. Campbell, Andrew J.; Danielson, Lisa; Righter, Kevin; Seagle, Christopher T.; Wang, Yanbin; Prakapenka, Vitali B. (2009). "High pressure effects on the iron–iron oxide and nickel–nickel oxide oxygen fugacity buffers". Earth and Planetary Science Letters. 286 (3–4): 556–564. Bibcode:2009E&PSL.286..556C. doi:10.1016/j.epsl.2009.07.022. ISSN   0012-821X.
  10. Frost, Daniel J. (2008-06-01). "The Upper Mantle and Transition Zone". Elements. 4 (3): 171–176. doi:10.2113/GSELEMENTS.4.3.171. ISSN   1811-5209. S2CID   129527426.
  11. McCammon, C.; Kopylova, M. G. (2004-07-17). "A redox profile of the Slave mantle and oxygen fugacity control in the cratonic mantle". Contributions to Mineralogy and Petrology. 148 (1): 55–68. Bibcode:2004CoMP..148...55M. doi:10.1007/s00410-004-0583-1. ISSN   0010-7999. S2CID   54778255.
  12. 1 2 3 Kiseeva, Ekaterina S.; Vasiukov, Denis M.; Wood, Bernard J.; McCammon, Catherine; Stachel, Thomas; Bykov, Maxim; Bykova, Elena; Chumakov, Aleksandr; Cerantola, Valerio (2018-01-22). "Oxidized iron in garnets from the mantle transition zone". Nature Geoscience. 11 (2): 144–147. Bibcode:2018NatGe..11..144K. doi:10.1038/s41561-017-0055-7. ISSN   1752-0894. S2CID   23720021.
  13. Rubie, David C.; Trønnes, Reidar G.; Catherine A. McCammon; Langenhorst, Falko; Liebske, Christian; Frost, Daniel J. (2004). "Experimental evidence for the existence of iron-rich metal in the Earth's lower mantle". Nature. 428 (6981): 409–412. Bibcode:2004Natur.428..409F. doi:10.1038/nature02413. ISSN   1476-4687. PMID   15042086. S2CID   32948214.
  14. Sverjensky, Dimitri A.; Huang, Fang (2015-11-03). "Diamond formation due to a pH drop during fluid–rock interactions". Nature Communications. 6 (1): 8702. Bibcode:2015NatCo...6.8702S. doi:10.1038/ncomms9702. ISSN   2041-1723. PMC   4667645 . PMID   26529259.
  15. Zhu, Feng; Li, Jie; Liu, Jiachao; Lai, Xiaojing; Chen, Bin; Meng, Yue (2019-02-18). "Kinetic Control on the Depth Distribution of Superdeep Diamonds". Geophysical Research Letters. 46 (4): 1984–1992. Bibcode:2019GeoRL..46.1984Z. doi: 10.1029/2018GL080740 . hdl: 2027.42/148362 .
  16. Pearson, D. G.; Brenker, F. E.; Nestola, F.; McNeill, J.; Nasdala, L.; Hutchison, M. T.; Matveev, S.; Mather, K.; Silversmit, G. (2014-03-12). "Hydrous mantle transition zone indicated by ringwoodite included within diamond" (PDF). Nature. 507 (7491): 221–224. Bibcode:2014Natur.507..221P. doi:10.1038/nature13080. ISSN   0028-0836. PMID   24622201. S2CID   205237822.
  17. Tschauner, O.; Huang, S.; Greenberg, E.; Prakapenka, V. B.; Ma, C.; Rossman, G. R.; Shen, A. H.; Zhang, D.; Newville, M. (2018-03-09). "Ice-VII inclusions in diamonds: Evidence for aqueous fluid in Earth's deep mantle". Science. 359 (6380): 1136–1139. Bibcode:2018Sci...359.1136T. doi: 10.1126/science.aao3030 . ISSN   0036-8075. PMID   29590042.
  18. Navon, Oded; Wirth, Richard; Schmidt, Christian; Jablon, Brooke Matat; Schreiber, Anja; Emmanuel, Simon (2017). "Solid molecular nitrogen ( δ -N 2 ) inclusions in Juina diamonds: Exsolution at the base of the transition zone". Earth and Planetary Science Letters. 464: 237–247. Bibcode:2017E&PSL.464..237N. doi:10.1016/j.epsl.2017.01.035.
  19. Nestola, F.; Korolev, N.; Kopylova, M.; Rotiroti, N.; Pearson, D. G.; Pamato, M. G.; Alvaro, M.; Peruzzo, L.; Gurney, J. J. (2018-03-07). "CaSiO3 perovskite in diamond indicates the recycling of oceanic crust into the lower mantle". Nature. 555 (7695): 237–241. Bibcode:2018Natur.555..237N. doi:10.1038/nature25972. ISSN   0028-0836. PMID   29516998. S2CID   3763653.