Reverse weathering

Last updated

Reverse weathering generally refers to the formation of a clay neoformation that utilizes cations and alkalinity in a process unrelated to the weathering of silicates. More specifically reverse weathering refers to the formation of authigenic clay minerals from the reaction of 1) biogenic silica with aqueous cations or cation bearing oxides or 2) cation poor precursor clays with dissolved cations or cation bearing oxides. [1]

Contents

Formation of cation abundant authigenic silicate clays is thought to occur through the following simplified reaction:

Biogenic opal (SiO2) + metal hydroxides (Al(OH)4) + dissolved cations (K+, Mg2+, Li+, etc.) + bicarbonate (HCO3) → clay minerals + H2O + CO2 [2]

The formation of authigenic clay minerals by reverse weathering is not fully understood. Much of the research done has been conducted in localized areas, such as the Amazon Delta, [3] Mississippi Delta, a palaeo-delta in Aínsa-Sobrarbe (Pyrenees) and in the Ethiopian Rift lakes, [4] making a global understanding of the process difficult. Much of the driving force behind research into reverse weathering stems from constraining the chemical mass balance between rivers and oceans. [5] Prior to the discovery of reverse weathering, the model of the chemical mass balance of the ocean predicted higher alkali metal and bicarbonate (HCO3) concentrations than was observed. [5] The formation of authigenic clay minerals was initially thought to account for the entirety of this excess, but the discovery of hydrothermal vents challenged this, as removal of alkali-alkaline earth metals and HCO3 from the ocean occurs in these locations as well. [5]

Methods of analysis

The process and extent of reverse weathering has been inferred by several methods and proxies.

In-situ measurements of biogenic silica and silicic acid (a product of weathering) have been used to analyze the rate and extent that reverse weathering occurs within an aquatic system. [6] [7] Uptake of biogenic silica as a result of reverse weathering would be observed as a relative low concentration of dissolved SiO2 compared to the overlying water.

Laboratory observations of reverse weathering have been conducted using incubations and flow through reactors to measure opal dissolution rates [2] [3] The clay was studied using scanning electron microscopes, x-ray, and transmission electron microscopes. [1] It was observed that the clay formed quickly, and using this amount of time and the known content of the sediment, concentration of potassium ions consumed by this process in rivers around the globe was estimated. [1]

Laboratory experiments can also include incubation experiments, in which sediment samples obtained from natural environments are enclosed in sealable containers with varied concentrations reverse weathering reactants (biogenic silica in the form of diatoms, cations, metals, etc.). [2]

ICP-OES Instrument Spectrometer ICP-OES.jpg
ICP-OES Instrument

Using an inductively coupled plasma optical emission spectrometer (ICP-OES) also provides concentration and isotopic information for cation and silica concentrations in pore water and digested sediment samples. Utilization of a multi-collector inductively coupled plasma mass spectrometer (MC-ICP-MS) is also used as a means of obtaining isotopic data of metals and silica in solution. [7]

Lithium isotope concentration within planktonic foraminifera has been used to infer past changes in silicate and reverse weathering rates over the last 68 million years. [8] Removal of lithium from seawater is mainly dependent on its assimilation within marine sediments and variations are believed to be indicative of the relative rates of silicate weathering and reverse weathering, in addition to other factors. Foraminifera with low lithium content suggest that reverse weathering may have been more prominent during that time period. [8]

Controls

Thermodynamics

Formation of authigenic silicate clays through reverse weathering was shown to be thermodynamically favorable during studies of Amazon delta sediments. [3] Primary controls on the formation of authigenic silicate clays are on the supply of reactants in solution. Areas of limited biogenic opal, metal hydroxides (e.g. aluminate (Al(OH)4)), or dissolved cations limit production of authigenic silicate clays. [2] Metals, cations, and silica are largely supplied by the weathering of terrigenous materials, which influences the thermodynamic favorability of reverse weathering. [9]

Kinetics

Kinetically, formation of clay minerals by reverse weathering can be relatively rapid (<1 year). [3] Due to the short formation timescale, reverse weathering is seen as a reasonable contributor to various ocean biogeochemical cycles. [3]

Influence on global cycles

Carbon cycle Carbon cycle-cute diagram.svg
Carbon cycle

The carbon cycle

The process of creating authigenic clay minerals through reverse weathering releases carbon dioxide (CO2). [9] However, release of bicarbonate by silicate weathering exceeds the quantities of CO2 produced by reverse weathering. Therefore, while reverse weathering does increase CO2 during production of authigenic clay minerals, it is overwhelmed by the concentration of HCO3 in the system, and will not have a significant effect on local pH. [9]

Basic diagram of the marine silica cycle, adapted from Treguer et al., 1995 Marine Silica Cycle.png
Basic diagram of the marine silica cycle, adapted from Treguer et al., 1995

The silica cycle

In recent years, the effect of reverse weathering on biogenic silica has been of great interest in quantifying the silica cycle. During weathering, dissolved silica is delivered to oceans through glacial runoff and riverine inputs. [2] This dissolved silica is taken up by a multitude of marine organisms, such as diatoms, and is used to create protective shells. [2] When these organisms die, they sink through the water column. [2] Without active production of biogenic SiO2, the mineral begins diagenesis. [2] Conversion of this dissolved silica into authigenic silicate clays through the process of reverse weathering constitutes a removal of 20-25% of silicon input. [11]

Reverse weathering is often found to occur in river deltas as these systems have high sediment accumulation rates and are observed to undergo rapid diagenesis. [12] The formation of silicate clays removes reactive silica from the pore waters of sediment, increasing the concentration of silica found in the rocks that form in these locations. [12]

Silicate weathering also appears to be a dominant process in deeper methanogenic sediments, whereas reverse weathering is more common in surface sediments, but still occurs at a lower rate. [3]

Study locations

The Amazon River Delta Amazon-river-delta-NASA.jpg
The Amazon River Delta

Deltas

In the Amazon River delta, about 90% of buried SiO2 is used up during reverse weathering, while the creation of potassium ions in this location is about 2.8 μmol g−1 year−1. [3] Nearly 7-10% of the potassium input from the Amazon River is removed from the ocean by the formation of potassium-iron rich aluminosilicates. [3] In the Mississippi River delta about 40% of SiO2 that is buried in the sediment is converted to authigenic aluminosilicates. [13] The major difference in the two deltas is due to the sediments in the Amazon delta being subject to a number of erosional and depositional processes, which creates an abundant amount of iron oxides. Sediment typically resides in the region for 30 years on average, but the upper layer undergoes major physical reworking 1-2 times per year. Pore water data suggests that the formation of authigenic clays in the Amazon delta occur on the order of a few months to a few years. The limiting reactant of clay formation in this region is the quantity of available SiO2, since the river water generally has a high concentration of other reactants, such as iron, potassium, magnesium, and aluminium. [3] Whereas in the Mississippi delta, the limiting nutrient for these reactions is iron.

The effect of reverse weathering has also been observed in paleo-delta systems. In the Ainsa basin, a palaeo-deltaic system was formed during the Eocene and uplifted through the orogeny of the Pyrenees. Isotopic geochemical differences were observed between palaeo sediments deposited in the marine conditions and those from alluvial environments. [14] The lithium isotope signature (δ7Li) and the silicon isotope signature (δ30Si) are systematically lighter in marine sediments than that in alluvial sediments, [14] implying authigenic clay formation in the marine sediments. Additionally, in the marine sediments the δ7Li signature is correlated to iron contents, suggesting the coupling of iron diagenesis and reverse weathering processes in the marine environments. This coupling can be achieved in reduced environments through the following reactions: [14]

H4SiO4 + Fe2+ + 2HCO3- → FeO-SiO2 (Fe-rich clays) + 3H2O + 2CO2

Ethiopian Rift Valley Lakes Abijatta - Langano - Shala.jpg
Ethiopian Rift Valley Lakes

Ethiopian rift lakes

Reverse weathering in the Ethiopian Rift lakes is easily observable, and recent studies at this location have been used to make inference on the extent of the process in the ocean. One study suggests that there is a general alkalinity deficit in the lakes, and that a little over half of this deficit can be attributed to the formation of aluminosilicate minerals. [4] The precipitation of salts is not profound, making the development of these clay minerals by reverse weathering more readily observable in comparison to the ocean. Using clay formation rates in the Ethiopian Rift lakes as a basis, the study suggests that clay formation in the ocean is too high to entirely attribute to the process of reverse weathering. It is believed that the deep-sea reverse weathering process never reaches completion, as pH is generally low. Hydrothermal activity is suggested to be a major contributor to clay formation in the deep ocean. [4]

Hydrothermal vents

Hydrothermal vent in the East Scotia Ridge East Scotia Ridge - Plos Biol 04.tif
Hydrothermal vent in the East Scotia Ridge

Some hypothesize that hydrothermal vents may be a prominent source of reverse weathering. [12] For some time, it was posited that terrestrial fluvial input was the only source of dissolved salts for the ocean. Later it was found that hydrothermal vents play a key role in the salinity of the oceans by releasing torrents of dissolved minerals that come from water/rock interactions. [15] At these locations, some dissolved salts react with rock and are removed, thus changing the ion composition of the seawater in comparison to the hydrothermal fluid. [15]

Some researchers hypothesize that reverse weathering could play a role in the silica cycle at hydrothermal vents. [5] Low temperature hydrothermal vents release silicic acid from the Earth's crust, and before it is able to exit the seabed, it cools and precipitates out as clay, such as a smectite. [11] The extent to which reverse weathering at hydrothermal vents adds to the overall silica cycle is a hot topic. [16] [17] [11]

History

In 1933, Victor Moritz Goldschmidt first proposed a reaction where igneous rock and volatiles would interact to generate sediments and seawater. [18] [19] [20] Lars Gunnar Sillén would later propose that reactions involving the generation of silicates potentially played a role in controlling the composition and pH of seawater in 1959. [19] At the time of Sillén's proposal, the thermodynamic constants of clay mineral reactions were not known and there were very few thermodynamic indicators that such a reaction existed. [21] Frederick Mackenzie and Robert Garrels would then combine Goldschmidt's and Sillén's work with the concept of reconstitution reactions to derive the reverse weathering hypothesis in 1966. [4] Since then, reverse weathering has been used as a possible explanation for various marine environment reactions and observations.

Today, there is much debate over the significance of reverse weathering. The global extent of the process has not yet been measured, but inferences can be made by using specific local examples. [22]

Related Research Articles

<span class="mw-page-title-main">Kaolinite</span> Layered non-swelling aluminosilicate 1:1 clay mineral

Kaolinite ( KAY-ə-lə-nete, -⁠lih-; also called kaolin) is a clay mineral, with the chemical composition Al2Si2O5(OH)4. It is a layered silicate mineral, with one tetrahedral sheet of silica (SiO4) linked through oxygen atoms to one octahedral sheet of alumina (AlO6) octahedra.

Geochemistry is the science that uses the tools and principles of chemistry to explain the mechanisms behind major geological systems such as the Earth's crust and its oceans. The realm of geochemistry extends beyond the Earth, encompassing the entire Solar System, and has made important contributions to the understanding of a number of processes including mantle convection, the formation of planets and the origins of granite and basalt. It is an integrated field of chemistry and geology.

<span class="mw-page-title-main">Shale</span> Fine-grained, clastic sedimentary rock

Shale is a fine-grained, clastic sedimentary rock formed from mud that is a mix of flakes of clay minerals (hydrous aluminium phyllosilicates, e.g. kaolin, Al2Si2O5(OH)4) and tiny fragments (silt-sized particles) of other minerals, especially quartz and calcite. Shale is characterized by its tendency to split into thin layers (laminae) less than one centimeter in thickness. This property is called fissility. Shale is the most common sedimentary rock.

<span class="mw-page-title-main">Chert</span> Hard, fine-grained sedimentary rock composed of cryptocrystalline silica

Chert is a hard, fine-grained sedimentary rock composed of microcrystalline or cryptocrystalline quartz, the mineral form of silicon dioxide (SiO2). Chert is characteristically of biological origin, but may also occur inorganically as a chemical precipitate or a diagenetic replacement, as in petrified wood.

<span class="mw-page-title-main">Diagenesis</span> Physico-chemical changes in sediments occurring after their deposition

Diagenesis is the process that describes physical and chemical changes in sediments first caused by water-rock interactions, microbial activity, and compaction after their deposition. Increased pressure and temperature only start to play a role as sediments become buried much deeper in the Earth's crust. In the early stages, the transformation of poorly consolidated sediments into sedimentary rock (lithification) is simply accompanied by a reduction in porosity and water expulsion, while their main mineralogical assemblages remain unaltered. As the rock is carried deeper by further deposition above, its organic content is progressively transformed into kerogens and bitumens.

<span class="mw-page-title-main">Smectite</span> Swelling 2:1 (TOT) phyllosilicates

A smectite is a mineral mixtures of various swelling sheet silicates (phyllosilicates), which have a three-layer 2:1 (TOT) structure and belong to the clay minerals. Smectites mainly consist of montmorillonite, but can often contain secondary minerals such as quartz and calcite.

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

The 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">Red beds</span> Sedimentary rocks with ferric oxides

Red beds are sedimentary rocks, typically consisting of sandstone, siltstone, and shale, that are predominantly red in color due to the presence of ferric oxides. Frequently, these red-colored sedimentary strata locally contain thin beds of conglomerate, marl, limestone, or some combination of these sedimentary rocks. The ferric oxides, which are responsible for the red color of red beds, typically occur as a coating on the grains of sediments comprising red beds. Classic examples of red beds are the Permian and Triassic strata of the western United States and the Devonian Old Red Sandstone facies of Europe.

Lithogenic silica (LSi) is silica (SiO2) derived from terrigenous rock (Igneous, metamorphic, and sedimentary), lithogenic sediments composed of the detritus of pre-existing rock, volcanic ejecta, extraterrestrial material, and minerals such silicate. Silica is the most abundant compound in the Earth's crust (59%) and the main component of almost every rock (>95%).

<span class="mw-page-title-main">Biogenic silica</span> Type of biogenic mineral

Biogenic silica (bSi), also referred to as opal, biogenic opal, or amorphous opaline silica, forms one of the most widespread biogenic minerals. For example, microscopic particles of silica called phytoliths can be found in grasses and other plants.

<span class="mw-page-title-main">Clastic rock</span> Sedimentary rocks made of mineral or rock fragments

Clastic rocks are composed of fragments, or clasts, of pre-existing minerals and rock. A clast is a fragment of geological detritus, chunks, and smaller grains of rock broken off other rocks by physical weathering. Geologists use the term clastic to refer to sedimentary rocks and particles in sediment transport, whether in suspension or as bed load, and in sediment deposits.

<span class="mw-page-title-main">Marine sediment</span>

Marine sediment, or ocean sediment, or seafloor sediment, are deposits of insoluble particles that have accumulated on the seafloor. These particles have their origins in soil and rocks and have been transported from the land to the sea, mainly by rivers but also by dust carried by wind and by the flow of glaciers into the sea. Additional deposits come from marine organisms and chemical precipitation in seawater, as well as from underwater volcanoes and meteorite debris.

<span class="mw-page-title-main">Siliceous ooze</span> Biogenic pelagic sediment located on the deep ocean floor

Siliceous ooze is a type of biogenic pelagic sediment located on the deep ocean floor. Siliceous oozes are the least common of the deep sea sediments, and make up approximately 15% of the ocean floor. Oozes are defined as sediments which contain at least 30% skeletal remains of pelagic microorganisms. Siliceous oozes are largely composed of the silica based skeletons of microscopic marine organisms such as diatoms and radiolarians. Other components of siliceous oozes near continental margins may include terrestrially derived silica particles and sponge spicules. Siliceous oozes are composed of skeletons made from opal silica SiO2·nH2O, as opposed to calcareous oozes, which are made from skeletons of calcium carbonate (CaCO3·nH2O) organisms (i.e. coccolithophores). Silica (Si) is a bioessential element and is efficiently recycled in the marine environment through the silica cycle. Distance from land masses, water depth and ocean fertility are all factors that affect the opal silica content in seawater and the presence of siliceous oozes.

<span class="mw-page-title-main">Carbonate–silicate cycle</span> Geochemical transformation of silicate rocks

The carbonate–silicate geochemical cycle, also known as the inorganic carbon cycle, describes the long-term transformation of silicate rocks to carbonate rocks by weathering and sedimentation, and the transformation of carbonate rocks back into silicate rocks by metamorphism and volcanism. Carbon dioxide is removed from the atmosphere during burial of weathered minerals and returned to the atmosphere through volcanism. On million-year time scales, the carbonate-silicate cycle is a key factor in controlling Earth's climate because it regulates carbon dioxide levels and therefore global temperature.

Carbonate-associated sulfates (CAS) are sulfate species found in association with carbonate minerals, either as inclusions, adsorbed phases, or in distorted sites within the carbonate mineral lattice. It is derived primarily from dissolved sulfate in the solution from which the carbonate precipitates. In the ocean, the source of this sulfate is a combination of riverine and atmospheric inputs, as well as the products of marine hydrothermal reactions and biomass remineralisation. CAS is a common component of most carbonate rocks, having concentrations in the parts per thousand within biogenic carbonates and parts per million within abiogenic carbonates. Through its abundance and sulfur isotope composition, it provides a valuable record of the global sulfur cycle across time and space.

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

The silica cycle is the biogeochemical cycle in which biogenic silica is transported between the Earth's systems. Silicon is considered a bioessential element and is one of the most abundant elements on Earth. The silica cycle has significant overlap with the carbon cycle and plays an important role in the sequestration of carbon through continental weathering, biogenic export and burial as oozes on geologic timescales.

Trace metal stable isotope biogeochemistry is the study of the distribution and relative abundances of trace metal isotopes in order to better understand the biological, geological, and chemical processes occurring in an environment. Trace metals are elements such as iron, magnesium, copper, and zinc that occur at low levels in the environment. Trace metals are critically important in biology and are involved in many processes that allow organisms to grow and generate energy. In addition, trace metals are constituents of numerous rocks and minerals, thus serving as an important component of the geosphere. Both stable and radioactive isotopes of trace metals exist, but this article focuses on those that are stable. Isotopic variations of trace metals in samples are used as isotopic fingerprints to elucidate the processes occurring in an environment and answer questions relating to biology, geochemistry, and medicine.

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

The lithium cycle (Li) is the biogeochemical cycle of lithium through the lithosphere and hydrosphere.

<span class="mw-page-title-main">Silicification</span> Geological petrification process

In geology, silicification is a petrification process in which silica-rich fluids seep into the voids of Earth materials, e.g., rocks, wood, bones, shells, and replace the original materials with silica (SiO2). Silica is a naturally existing and abundant compound found in organic and inorganic materials, including Earth's crust and mantle. There are a variety of silicification mechanisms. In silicification of wood, silica permeates into and occupies cracks and voids in wood such as vessels and cell walls. The original organic matter is retained throughout the process and will gradually decay through time. In the silicification of carbonates, silica replaces carbonates by the same volume. Replacement is accomplished through the dissolution of original rock minerals and the precipitation of silica. This leads to a removal of original materials out of the system. Depending on the structures and composition of the original rock, silica might replace only specific mineral components of the rock. Silicic acid (H4SiO4) in the silica-enriched fluids forms lenticular, nodular, fibrous, or aggregated quartz, opal, or chalcedony that grows within the rock. Silicification happens when rocks or organic materials are in contact with silica-rich surface water, buried under sediments and susceptible to groundwater flow, or buried under volcanic ashes. Silicification is often associated with hydrothermal processes. Temperature for silicification ranges in various conditions: in burial or surface water conditions, temperature for silicification can be around 25°−50°; whereas temperatures for siliceous fluid inclusions can be up to 150°−190°. Silicification could occur during a syn-depositional or a post-depositional stage, commonly along layers marking changes in sedimentation such as unconformities or bedding planes.

<span class="mw-page-title-main">Silicon isotope biogeochemistry</span> The study of environmental processes using the relative abundance of Si isotopes

Silicon isotope biogeochemistry is the study of environmental processes using the relative abundance of Si isotopes. As the relative abundance of Si stable isotopes varies among different natural materials, the differences in abundance can be used to trace the source of Si, and to study biological, geological, and chemical processes. The study of stable isotope biogeochemistry of Si aims to quantify the different Si fluxes in the global biogeochemical silicon cycle, to understand the role of biogenic silica within the global Si cycle, and to investigate the applications and limitations of the sedimentary Si record as an environmental and palaeoceanographic proxy.

References

  1. 1 2 3 Mackenzie, Fred T.; Kump, Lee R. (1995-10-27). "Reverse Weathering, Clay Mineral Formation, and Oceanic Element Cycles". Science. 270 (5236): 586. Bibcode:1995Sci...270..586M. doi:10.1126/science.270.5236.586. ISSN   0036-8075. S2CID   128993379.
  2. 1 2 3 4 5 6 7 8 Loucaides, Socratis; Michalopoulos, Panagiotis; Presti, Massimo; Koning, Erica; Behrends, Thilo; Van Cappellen, Philippe (2010-02-15). "Seawater-mediated interactions between diatomaceous silica and terrigenous sediments: Results from long-term incubation experiments". Chemical Geology. 270 (1–4): 68–79. Bibcode:2010ChGeo.270...68L. doi:10.1016/j.chemgeo.2009.11.006.
  3. 1 2 3 4 5 6 7 8 9 Michalopoulos, Panagiotis; Aller, Robert C (2004-03-01). "Early diagenesis of biogenic silica in the Amazon delta: alteration, authigenic clay formation, and storage". Geochimica et Cosmochimica Acta. 68 (5): 1061–1085. Bibcode:2004GeCoA..68.1061M. doi:10.1016/j.gca.2003.07.018.
  4. 1 2 3 4 Damm, K. L.; Edmond, J. M. (1984). "Reverse Weathering in the Closed-Basin Lakes of the Ethiopian Rift". American Journal of Science. 284 (7): 835–862. doi:10.2475/ajs.284.7.835.
  5. 1 2 3 4 Mackenzie, Fred; Garrels, Robert (1966). "Chemical Mass Balance Between Rivers and Oceans". American Journal of Science. 264 (7): 507–525. Bibcode:1966AmJS..264..507M. doi: 10.2475/ajs.264.7.507 .
  6. Morifuji, Naoto; Nakashima, Satoru (2016). "Hydrothermal transformation of biogenic silica as studied by in situ infrared spectroscopy" (PDF). Goldscmidt Conference Abstracts.
  7. 1 2 Ehlert, Claudia; Doering, Kristin; Wallmann, Klaus; Scholz, Florian; Sommer, Stefan; Grasse, Patricia; Geilert, Sonja; Frank, Martin (2016). "Stable silicon isotope signatures of marine pore waters – Biogenic opal dissolution versus authigenic clay mineral formation". Geochimica et Cosmochimica Acta. 191: 102–117. Bibcode:2016GeCoA.191..102E. doi:10.1016/j.gca.2016.07.022.
  8. 1 2 Misra, Sambuddha; Froelich, Philip (2012). "Lithium Isotope History of Cenozoic Seawater: Changes in Silicate Weathering and Reverse Weathering". Science. 335 (6070): 818–823. Bibcode:2012Sci...335..818M. doi: 10.1126/science.1214697 . PMID   22282473. S2CID   42591236.
  9. 1 2 3 Li, Gaojun; Elderfield, Henry (2013-02-15). "Evolution of carbon cycle over the past 100 million years". Geochimica et Cosmochimica Acta. 103: 11–25. Bibcode:2013GeCoA.103...11L. doi:10.1016/j.gca.2012.10.014.
  10. Tréguer, Paul; Nelson, David M.; Bennekom, Aleido J. Van; DeMaster, David J.; Leynaert, Aude; Quéguiner, Bernard (1995-04-21). "The Silica Balance in the World Ocean: A Reestimate". Science. 268 (5209): 375–379. Bibcode:1995Sci...268..375T. doi:10.1126/science.268.5209.375. ISSN   0036-8075. PMID   17746543. S2CID   5672525.
  11. 1 2 3 Tréguer, Paul J.; Rocha, Christina L. De La (2013-01-02). "The World Ocean Silica Cycle". Annual Review of Marine Science. 5 (1): 477–501. doi:10.1146/annurev-marine-121211-172346. PMID   22809182.
  12. 1 2 3 Aller, R. C. (2014-01-01). "Sedimentary Diagenesis, Depositional Environments, and Benthic Fluxes". In Holland, Heinrich D.; Turekian, Karl K. (eds.). Treatise on Geochemistry (Second ed.). Oxford: Elsevier. pp. 293–334. doi:10.1016/b978-0-08-095975-7.00611-2. ISBN   9780080983004.
  13. Presti, Massimo; Michalopoulos, Panagiotis (2008-04-01). "Estimating the contribution of the authigenic mineral component to the long-term reactive silica accumulation on the western shelf of the Mississippi River Delta". Continental Shelf Research. 28 (6): 823–838. Bibcode:2008CSR....28..823P. doi:10.1016/j.csr.2007.12.015.
  14. 1 2 3 Zhang, Xu (Yvon); Gaillardet, Jérôme; Barrier, Laurie; Bouchez, Julien (2022-01-15). "Li and Si isotopes reveal authigenic clay formation in a palaeo-delta". Earth and Planetary Science Letters. 578: 117339. Bibcode:2022E&PSL.57817339Z. doi:10.1016/j.epsl.2021.117339. ISSN   0012-821X. S2CID   245366429.
  15. 1 2 Von Damm, K (1995). "Controls on the Chemistry and Temporal Variability of Seafloor Hydrothermal Fluids". Geophysical Monograph. Geophysical Monograph Series. 91: 222–247. Bibcode:1995GMS....91..222V. doi:10.1029/GM091p0222. ISBN   9781118663998 via American Geophysical Union.
  16. Rahman, S.; Aller, R. C.; Cochran, J. K. (2016-07-16). "Cosmogenic 32Si as a tracer of biogenic silica burial and diagenesis: Major deltaic sinks in the silica cycle". Geophysical Research Letters. 43 (13): 2016GL069929. Bibcode:2016GeoRL..43.7124R. doi: 10.1002/2016GL069929 . ISSN   1944-8007.
  17. Rasmussen, Birger (1998). "Removal of oceanic REE by authigenic precipitation of phosphatic minerals". Earth and Planetary Science Letters. 164 (1–2): 135–149. Bibcode:1998E&PSL.164..135R. doi:10.1016/S0012-821X(98)00199-X.
  18. Li, Yuan-Hui (1972). "Geochemical Mass Balance Among Lithosphere, Hydrosphere, and Atmosphere". American Journal of Science. 272 (2): 119–137. Bibcode:1972AmJS..272..119L. doi: 10.2475/ajs.272.2.119 .
  19. 1 2 Ristvet, Byron (1978). "Reverse Weathering Reactions Within Recent Nearshore Sediments, Kaneohe Bay, Oahu". Defense Nuclear Agency.
  20. Sillen, Lars (1967). "The Ocean as a Chemical System". Science. 156 (3779): 1189–1197. Bibcode:1967Sci...156.1189S. doi:10.1126/science.156.3779.1189. PMID   17792775. S2CID   220099254.
  21. Emerson, Steven; Hedges, John (2008). Chemical Oceanography and the Marine Carbon Cycle . US, Cambridge University Press, New York: Cambridge University Press. pp.  43–44. ISBN   978-0-521-83313-4.
  22. Holland, H.D.; Turekian, K.K., eds. (2014). "Sedimentary Diagenesis, Depositional Environments, and Benthic Fluxes". Treatise on Geochemistry. Vol. 8 (2 ed.). Oxford: Elsevier. pp. 293–334.