Microbial loop

Last updated
The aquatic microbial loop is a marine trophic pathway which incorporates dissolved organic carbon into the food chain. Microbial Loop.jpg
The aquatic microbial loop is a marine trophic pathway which incorporates dissolved organic carbon into the food chain.

The microbial loop describes a trophic pathway where, in aquatic systems, dissolved organic carbon (DOC) is returned to higher trophic levels via its incorporation into bacterial biomass, and then coupled with the classic food chain formed by phytoplankton-zooplankton-nekton. In soil systems, the microbial loop refers to soil carbon. The term microbial loop was coined by Farooq Azam, Tom Fenchel et al. [1] in 1983 to include the role played by bacteria in the carbon and nutrient cycles of the marine environment.

Contents

In general, dissolved organic carbon (DOC) is introduced into the ocean environment from bacterial lysis, the leakage or exudation of fixed carbon from phytoplankton (e.g., mucilaginous exopolymer from diatoms), sudden cell senescence, sloppy feeding by zooplankton, the excretion of waste products by aquatic animals, or the breakdown or dissolution of organic particles from terrestrial plants and soils. [2] Bacteria in the microbial loop decompose this particulate detritus to utilize this energy-rich matter for growth. Since more than 95% of organic matter in marine ecosystems consists of polymeric, high molecular weight (HMW) compounds (e.g., protein, polysaccharides, lipids), only a small portion of total dissolved organic matter (DOM) is readily utilizable to most marine organisms at higher trophic levels. This means that dissolved organic carbon is not available directly to most marine organisms; marine bacteria introduce this organic carbon into the food web, resulting in additional energy becoming available to higher trophic levels. Recently the term "microbial food web" has been substituted for the term "microbial loop".

History

Prior to the discovery of the microbial loop, the classic view of marine food webs was one of a linear chain from phytoplankton to nekton. Generally, marine bacteria were not thought to be significant consumers of organic matter (including carbon), although they were known to exist. However, the view of a marine pelagic food web was challenged during the 1970s and 1980s by Pomeroy and Azam, who suggested the alternative pathway of carbon flow from bacteria to protozoans to metazoans. [3] [1]

Early work in marine ecology that investigated the role of bacteria in oceanic environments concluded their role to be very minimal. Traditional methods of counting bacteria (e.g., culturing on agar plates) only yielded small numbers of bacteria that were much smaller than their true ambient abundance in seawater. Developments in technology for counting bacteria have led to an understanding of the significant importance of marine bacteria in oceanic environments.

In the 1970s, the alternative technique of direct microscopic counting was developed by Francisco et al. (1973) and Hobbie et al. (1977). Bacterial cells were counted with an epifluorescence microscope, producing what is called an "acridine orange direct count" (AODC). This led to a reassessment of the large concentration of bacteria in seawater, which was found to be more than was expected (typically on the order of 1 million per milliliter). Also, development of the "bacterial productivity assay" showed that a large fraction (i.e. 50%) of net primary production (NPP) was processed by marine bacteria.

In 1974, Larry Pomeroy published a paper in BioScience entitled "The Ocean's Food Web: A Changing Paradigm", where the key role of microbes in ocean productivity was highlighted. [3] In the early 1980s, Azam and a panel of top ocean scientists published the synthesis of their discussion in the journal Marine Ecology Progress Series entitled "The Ecological Role of Water Column Microbes in the Sea". The term 'microbial loop' was introduced in this paper, which noted that the bacteria-consuming protists were in the same size class as phytoplankton and likely an important component of the diet of planktonic crustaceans. [1]

Evidence accumulated since this time has indicated that some of these bacterivorous protists (such as ciliates) are actually selectively preyed upon by these copepods. In 1986, Prochlorococcus , which is found in high abundance in oligotrophic areas of the ocean, was discovered by Sallie W. Chisholm, Robert J. Olson, and other collaborators (although there had been several earlier records of very small cyanobacteria containing chlorophyll b in the ocean [4] [5] Prochlorococcus was discovered in 1986 [6] ). [7] Stemming from this discovery, researchers observed the changing role of marine bacteria along a nutrient gradient from eutrophic to oligotrophic areas in the ocean.

Factors controlling the microbial loop

The efficiency of the microbial loop is determined by the density of marine bacteria within it. [8] It has become clear that bacterial density is mainly controlled by the grazing activity of small protozoans and various taxonomic groups of flagellates. Also, viral infection causes bacterial lysis, which release cell contents back into the dissolved organic matter (DOM) pool, lowering the overall efficiency of the microbial loop. Mortality from viral infection has almost the same magnitude as that from protozoan grazing. However, compared to protozoan grazing, the effect of viral lysis can be very different because lysis is highly host-specific to each marine bacteria. Both protozoan grazing and viral infection balance the major fraction of bacterial growth. In addition, the microbial loop dominates in oligotrophic waters, rather than in eutrophic areas - there the classical plankton food chain predominates, due to the frequent fresh supply of mineral nutrients (e.g. spring bloom in temperate waters, upwelling areas). The magnitude of the efficiency of the microbial loop can be determined by measuring bacterial incorporation of radiolabeled substrates (such as tritiated thymidine or leucine).

In marine ecosystems

The microbial loop is of particular importance in increasing the efficiency of the marine food web via the utilization of dissolved organic matter (DOM), which is typically unavailable to most marine organisms. In this sense, the process aids in recycling of organic matter and nutrients and mediates the transfer of energy above the thermocline. More than 30% of dissolved organic carbon (DOC) incorporated into bacteria is respired and released as carbon dioxide. The other main effect of the microbial loop in the water column is that it accelerates mineralization through regenerating production in nutrient-limited environments (e.g. oligotrophic waters). In general, the entire microbial loop is to some extent typically five to ten times the mass of all multicellular marine organisms in the marine ecosystem. Marine bacteria are the base of the food web in most oceanic environments, and they improve the trophic efficiency of both marine food webs and important aquatic processes (such as the productivity of fisheries and the amount of carbon exported to the ocean floor). Therefore, the microbial loop, together with primary production, controls the productivity of marine systems in the ocean.

Many planktonic bacteria are motile, using a flagellum to propagate, and chemotax to locate, move toward, and attach to a point source of dissolved organic matter (DOM) where fast growing cells digest all or part of the particle. Accumulation within just a few minutes at such patches is directly observable. Therefore, the water column can be considered to some extent as a spatially organized place on a small scale rather than a completely mixed system. This patch formation affects the biologically-mediated transfer of matter and energy in the microbial loop.

More currently, the microbial loop is considered to be more extended. [9] Chemical compounds in typical bacteria (such as DNA, lipids, sugars, etc.) and similar values of C:N ratios per particle are found in the microparticles formed abiotically. Microparticles are a potentially attractive food source to bacterivorous plankton. If this is the case, the microbial loop can be extended by the pathway of direct transfer of dissolved organic matter (DOM) via abiotic microparticle formation to higher trophic levels. This has ecological importance in two ways. First, it occurs without carbon loss, and makes organic matter more efficiently available to phagotrophic organisms, rather than only heterotrophic bacteria. Furthermore, abiotic transformation in the extended microbial loop depends only on temperature and the capacity of DOM to aggregate, while biotic transformation is dependent on its biological availability. [9]

In land ecosystems

Soil carbon cycle through the microbial loop Soil carbon cycle through the microbial loop.jpg
Soil carbon cycle through the microbial loop

Soil ecosystems are highly complex and subject to different landscape-scale perturbations that govern whether soil carbon is retained or released to the atmosphere. [11] The ultimate fate of soil organic carbon is a function of the combined activities of plants and below ground organisms, including soil microbes. Although soil microorganisms are known to support a plethora of biogeochemical functions related to carbon cycling, [12] the vast majority of the soil microbiome remains uncultivated and has largely cryptic functions. [13] Only a mere fraction of soil microbial life has been catalogued to date, although new soil microbes [13] and viruses are increasingly being discovered. [14] This lack of knowledge results in uncertainty of the contribution of soil microorganisms to soil organic carbon cycling and hinders construction of accurate predictive models for global carbon flux under climate change. [15] [10]

The lack of information concerning the soil microbiome metabolic potential makes it particularly challenging to accurately account for the shifts in microbial activities that occur in response to environmental change. For example, plant-derived carbon inputs can prime microbial activity to decompose existing soil organic carbon at rates higher than model expectations, resulting in error within predictive models of carbon fluxes. [16] [10]

To account for this, a conceptual model known as the microbial carbon pump, illustrated in the diagram on the right, has been developed to define how soil microorganisms transform and stabilise soil organic matter. [17] As shown in the diagram, carbon dioxide in the atmosphere is fixed by plants (or autotrophic microorganisms) and added to soil through processes such as (1) root exudation of low-molecular weight simple carbon compounds, or deposition of leaf and root litter leading to accumulation of complex plant polysaccharides. (2) Through these processes, carbon is made bioavailable to the microbial metabolic "factory" and subsequently is either (3) respired to the atmosphere or (4) enters the stable carbon pool as microbial necromass. The exact balance of carbon efflux versus persistence is a function of several factors, including aboveground plant community composition and root exudate profiles, environmental variables, and collective microbial phenotypes (i.e., the metaphenome). [18] [10]

In this model, microbial metabolic activities for carbon turnover are segregated into two categories: ex vivo modification, referring to transformation of plant-derived carbon by extracellular enzymes, and in vivo turnover, for intracellular carbon used in microbial biomass turnover or deposited as dead microbial biomass, referred to as necromass. The contrasting impacts of catabolic activities that release soil organic carbon as carbon dioxide (CO2), versus anabolic pathways that produce stable carbon compounds, control net carbon retention rates. In particular, microbial carbon sequestration represents an underrepresented aspect of soil carbon flux that the microbial carbon pump model attempts to address. [17] A related area of uncertainty is how the type of plant-derived carbon enhances microbial soil organic carbon storage or alternatively accelerates soil organic carbon decomposition. [19] For example, leaf litter and needle litter serve as sources of carbon for microbial growth in forest soils, but litter chemistry and pH varies by vegetation type [e.g., between root and foliar litter [20] or between deciduous and coniferous forest litter (14)]. In turn, these biochemical differences influence soil organic carbon levels through changing decomposition dynamics. [21] Also, increased diversity of plant communities increases rates of rhizodeposition, stimulating microbial activity and soil organic carbon storage, [22] although soils eventually reach a saturation point beyond which they cannot store additional carbon. [23] [10]

See also

Related Research Articles

<span class="mw-page-title-main">Plankton</span> Organisms living in water or air that are drifters on the current or wind

Plankton are the diverse collection of organisms found in water that are unable to propel themselves against a current. The individual organisms constituting plankton are called plankters. In the ocean, they provide a crucial source of food to many small and large aquatic organisms, such as bivalves, fish, and baleen whales.

<span class="mw-page-title-main">Zooplankton</span> Heterotrophic protistan or metazoan members of the plankton ecosystem

Zooplankton are the animal component of the planktonic community, having to consume other organisms to thrive. Plankton are aquatic organisms that are unable to swim effectively against currents. Consequently, they drift or are carried along by currents in the ocean, or by currents in seas, lakes or rivers.

<span class="mw-page-title-main">Biological pump</span> Carbon capture process in oceans

The biological pump (or ocean carbon biological pump or marine biological carbon pump) is the ocean's biologically driven sequestration of carbon from the atmosphere and land runoff to the ocean interior and seafloor sediments. In other words, it is a biologically mediated process which results in the sequestering of carbon in the deep ocean away from the atmosphere and the land. The biological pump is the biological component of the "marine carbon pump" which contains both a physical and biological component. It is the part of the broader oceanic carbon cycle responsible for the cycling of organic matter formed mainly by phytoplankton during photosynthesis (soft-tissue pump), as well as the cycling of calcium carbonate (CaCO3) formed into shells by certain organisms such as plankton and mollusks (carbonate pump).

Photoheterotrophs are heterotrophic phototrophs—that is, they are organisms that use light for energy, but cannot use carbon dioxide as their sole carbon source. Consequently, they use organic compounds from the environment to satisfy their carbon requirements; these compounds include carbohydrates, fatty acids, and alcohols. Examples of photoheterotrophic organisms include purple non-sulfur bacteria, green non-sulfur bacteria, and heliobacteria. These microorganisms are ubiquitous in aquatic habitats, occupy unique niche-spaces, and contribute to global biogeochemical cycling. Recent research has also indicated that the oriental hornet and some aphids may be able to use light to supplement their energy supply.

<span class="mw-page-title-main">Dissolved organic carbon</span> Organic carbon classification

Dissolved organic carbon (DOC) is the fraction of organic carbon operationally defined as that which can pass through a filter with a pore size typically between 0.22 and 0.7 micrometers. The fraction remaining on the filter is called particulate organic carbon (POC).

<span class="mw-page-title-main">Picoplankton</span> Fraction of plankton between 0.2 and 2 μm

Picoplankton is the fraction of plankton composed by cells between 0.2 and 2 μm that can be either prokaryotic and eukaryotic phototrophs and heterotrophs:

The microbial food web refers to the combined trophic interactions among microbes in aquatic environments. These microbes include viruses, bacteria, algae, heterotrophic protists.

<span class="mw-page-title-main">Sea surface microlayer</span> Boundary layer where all exchange occurs between the atmosphere and the ocean

The sea surface microlayer (SML) is the boundary interface between the atmosphere and ocean, covering about 70% of Earth's surface. With an operationally defined thickness between 1 and 1,000 μm (1.0 mm), the SML has physicochemical and biological properties that are measurably distinct from underlying waters. Recent studies now indicate that the SML covers the ocean to a significant extent, and evidence shows that it is an aggregate-enriched biofilm environment with distinct microbial communities. Because of its unique position at the air-sea interface, the SML is central to a range of global marine biogeochemical and climate-related processes.

<span class="mw-page-title-main">Marine snow</span> Shower of organic detritus in the ocean

In the deep ocean, marine snow is a continuous shower of mostly organic detritus falling from the upper layers of the water column. It is a significant means of exporting energy from the light-rich photic zone to the aphotic zone below, which is referred to as the biological pump. Export production is the amount of organic matter produced in the ocean by primary production that is not recycled (remineralised) before it sinks into the aphotic zone. Because of the role of export production in the ocean's biological pump, it is typically measured in units of carbon. The term was coined by explorer William Beebe as observed from his bathysphere. As the origin of marine snow lies in activities within the productive photic zone, the prevalence of marine snow changes with seasonal fluctuations in photosynthetic activity and ocean currents. Marine snow can be an important food source for organisms living in the aphotic zone, particularly for organisms that live very deep in the water column.

<span class="mw-page-title-main">Bacterioplankton</span> Bacterial component of the plankton that drifts in the water column

Bacterioplankton refers to the bacterial component of the plankton that drifts in the water column. The name comes from the Ancient Greek word πλανκτος, meaning "wanderer" or "drifter", and bacterium, a Latin term coined in the 19th century by Christian Gottfried Ehrenberg. They are found in both seawater and freshwater.

<span class="mw-page-title-main">Marine microorganisms</span> Any life form too small for the naked human eye to see that lives in a marine environment

Marine microorganisms are defined by their habitat as microorganisms living in a marine environment, that is, in the saltwater of a sea or ocean or the brackish water of a coastal estuary. A microorganism is any microscopic living organism or virus, that is too small to see with the unaided human eye without magnification. Microorganisms are very diverse. They can be single-celled or multicellular and include bacteria, archaea, viruses and most protozoa, as well as some fungi, algae, and animals, such as rotifers and copepods. Many macroscopic animals and plants have microscopic juvenile stages. Some microbiologists also classify viruses as microorganisms, but others consider these as non-living.

<span class="mw-page-title-main">Plastisphere</span> Plastic debris suspended in water and organisms which live in it

The plastisphere consists of ecosystems that have evolved to live in human-made plastic environments. All plastic accumulated in marine ecosystems serves as a habitat for various types of microorganisms, with the most notable contaminant being microplastics. There are an estimate of about 51 trillion microplastics floating in the oceans. Relating to the plastisphere, over 1,000 different species of microbes are able to inhabit just one of these 5mm pieces of plastic.

<span class="mw-page-title-main">Phycosphere</span> Microscale mucus region that is rich in organic matter surrounding a phytoplankton cel

The phycosphere is a microscale mucus region that is rich in organic matter surrounding a phytoplankton cell. This area is high in nutrients due to extracellular waste from the phytoplankton cell and it has been suggested that bacteria inhabit this area to feed on these nutrients. This high nutrient environment creates a microbiome and a diverse food web for microbes such as bacteria and protists. It has also been suggested that the bacterial assemblages within the phycosphere are species-specific and can vary depending on different environmental factors.

<span class="mw-page-title-main">Mycoplankton</span> Fungal members of the plankton communities of aquatic ecosystems

Mycoplankton are saprotrophic members of the plankton communities of marine and freshwater ecosystems. They are composed of filamentous free-living fungi and yeasts that are associated with planktonic particles or phytoplankton. Similar to bacterioplankton, these aquatic fungi play a significant role in heterotrophicmineralization and nutrient cycling. Mycoplankton can be up to 20 mm in diameter and over 50 mm in length.

<span class="mw-page-title-main">Particulate organic matter</span>

Particulate organic matter (POM) is a fraction of total organic matter operationally defined as that which does not pass through a filter pore size that typically ranges in size from 0.053 millimeters (53 μm) to 2 millimeters.

Mary Ann Moran is a distinguished research professor of marine sciences at the University of Georgia in Athens. She studies the role of bacteria in Earth's marine nutrient cycles, and is a leader in the fields of marine sciences and biogeochemistry. Her work is focused on how microbes interact with dissolved organic matter and the impact of microbial diversity on the global carbon and sulfur cycles. By defining the roles of diverse bacteria in the carbon and sulfur cycles, she connects the biogeochemical and organismal approaches in marine science.

<span class="mw-page-title-main">Sea ice microbial communities</span> Groups of microorganisms living within and at the interfaces of sea ice

Sea Ice Microbial Communities (SIMCO) refer to groups of microorganisms living within and at the interfaces of sea ice at the poles. The ice matrix they inhabit has strong vertical gradients of salinity, light, temperature and nutrients. Sea ice chemistry is most influenced by the salinity of the brine which affects the pH and the concentration of dissolved nutrients and gases. The brine formed during the melting sea ice creates pores and channels in the sea ice in which these microbes can live. As a result of these gradients and dynamic conditions, a higher abundance of microbes are found in the lower layer of the ice, although some are found in the middle and upper layers. Despite this extreme variability in environmental conditions, the taxonomical community composition tends to remain consistent throughout the year, until the ice melts.

<span class="mw-page-title-main">Viral shunt</span>

The viral shunt is a mechanism that prevents marine microbial particulate organic matter (POM) from migrating up trophic levels by recycling them into dissolved organic matter (DOM), which can be readily taken up by microorganisms. The DOM recycled by the viral shunt pathway is comparable to the amount generated by the other main sources of marine DOM.

<span class="mw-page-title-main">Marine food web</span> Marine consumer-resource system

Compared to terrestrial environments, marine environments have biomass pyramids which are inverted at the base. In particular, the biomass of consumers is larger than the biomass of primary producers. This happens because the ocean's primary producers are tiny phytoplankton which grow and reproduce rapidly, so a small mass can have a fast rate of primary production. In contrast, many significant terrestrial primary producers, such as mature forests, grow and reproduce slowly, so a much larger mass is needed to achieve the same rate of primary production.

Cytophagales is an order of non-spore forming, rod-shaped, Gram-negative bacteria that move through a gliding or flexing motion. These chemoorganotrophs are important remineralizers of organic materials into micronutrients. They are widely dispersed in the environment, found in ecosystems including soil, freshwater, seawater and sea ice. Cytophagales is included in the Bacteroidota phylum.

References

  1. 1 2 3 Azam, Farooq; Fenchel, Tom; Field, J.G.; Gray, J.S.; Meyer-Reil, L.A.; Thingstad, F. (1983). "The Ecological Role of Water-Column Microbes in the Sea". Marine Ecology Progress Series. 10: 257–263. Bibcode:1983MEPS...10..257A. doi: 10.3354/meps010257 .
  2. Van den Meersche, Karel; Middelburg, Jack J.; Soetaert, Karline; van Rijswijk, Pieter; Boschker, Henricus T. S.; Heip, Carlo H. R. (2004). "Carbon-nitrogen coupling and algal-bacterial interactions during an experimental bloom: Modeling a13C tracer experiment". Limnology and Oceanography. 49 (3): 862–878. Bibcode:2004LimOc..49..862V. doi:10.4319/lo.2004.49.3.0862. hdl: 1854/LU-434810 . ISSN   0024-3590.
  3. 1 2 Pomeroy, Lawrence R. (1974). "The Ocean's Food Web, A Changing Paradigm". BioScience. 24 (9): 499–504. doi:10.2307/1296885. ISSN   0006-3568. JSTOR   1296885.
  4. Johnson, P. W.; Sieburth, J. M. (1979). "Chroococcoid cyanobacteria in the sea: a ubiquitous and diverse phototrophic biomass". Limnology and Oceanography. 24 (5): 928–935. Bibcode:1979LimOc..24..928J. doi:10.4319/lo.1979.24.5.0928.
  5. Gieskes, W. W. C.; Kraay, G. W. (1983). "Unknown chlorophyll a derivatives in the North Sea and the tropical Atlantic Ocean revealed by HPLC analysis". Limnology and Oceanography. 28 (4): 757–766. Bibcode:1983LimOc..28..757G. doi: 10.4319/lo.1983.28.4.0757 .
  6. Chisholm, S. W.; Olson, R. J.; Zettler, E. R.; Waterbury, J.; Goericke, R.; Welschmeyer, N. (1988). "A novel free-living prochlorophyte occurs at high cell concentrations in the oceanic euphotic zone". Nature . 334 (6180): 340–343. Bibcode:1988Natur.334..340C. doi:10.1038/334340a0. S2CID   4373102.
  7. Chisholm, Sallie W.; Frankel, Sheila L.; Goericke, Ralf; Olson, Robert J.; Palenik, Brian; Waterbury, John B.; West-Johnsrud, Lisa; Zettler, Erik R. (1992). "Prochlorococcus marinus nov. gen. nov. sp.: an oxyphototrophic marine prokaryote containing divinyl chlorophyll a and b". Archives of Microbiology. 157 (3): 297–300. doi:10.1007/bf00245165. ISSN   0302-8933. S2CID   32682912.
  8. Taylor, AH; Joint, I (1990). "A steady-state analysis of the 'microbial loop' in stratified systems". Marine Ecology Progress Series. Inter-Research Science Center. 59: 1–17. Bibcode:1990MEPS...59....1T. doi: 10.3354/meps059001 . ISSN   0171-8630.
  9. 1 2 Kerner, Martin; Hohenberg, Heinz; Ertl, Siegmund; Reckermann, Marcus; Spitzy, Alejandro (2003). "Self-organization of dissolved organic matter to micelle-like microparticles in river water". Nature. Springer Science and Business Media LLC. 422 (6928): 150–154. Bibcode:2003Natur.422..150K. doi:10.1038/nature01469. ISSN   0028-0836. PMID   12634782. S2CID   4380194.
  10. 1 2 3 4 5 Naylor, Dan; Sadler, Natalie; Bhattacharjee, Arunima; Graham, Emily B.; Anderton, Christopher R.; McClure, Ryan; Lipton, Mary; Hofmockel, Kirsten S.; Jansson, Janet K. (2020). "Soil Microbiomes Under Climate Change and Implications for Carbon Cycling". Annual Review of Environment and Resources. 45: 29–59. doi: 10.1146/annurev-environ-012320-082720 . CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  11. Jansson, Janet K.; Hofmockel, Kirsten S. (2020). "Soil microbiomes and climate change". Nature Reviews Microbiology. 18 (1): 35–46. doi:10.1038/s41579-019-0265-7. OSTI   1615017. PMID   31586158. S2CID   203658781.
  12. Thompson, Luke R.; et al. (2017). "A communal catalogue reveals Earth's multiscale microbial diversity". Nature. 551 (7681): 457–463. Bibcode:2017Natur.551..457T. doi:10.1038/nature24621. PMC   6192678 . PMID   29088705.
  13. 1 2 Goel, Reeta; Kumar, Vinay; Suyal, Deep Chandra; Narayan; Soni, Ravindra (2018). "Toward the Unculturable Microbes for Sustainable Agricultural Production". Role of Rhizospheric Microbes in Soil. pp. 107–123. doi:10.1007/978-981-10-8402-7_4. ISBN   978-981-10-8401-0.
  14. Schulz, Frederik; Alteio, Lauren; Goudeau, Danielle; Ryan, Elizabeth M.; Yu, Feiqiao B.; Malmstrom, Rex R.; Blanchard, Jeffrey; Woyke, Tanja (2018). "Hidden diversity of soil giant viruses". Nature Communications. 9 (1): 4881. Bibcode:2018NatCo...9.4881S. doi:10.1038/s41467-018-07335-2. PMC   6243002 . PMID   30451857.
  15. Zeng, N.; Yoshikawa, C.; Weaver, A. J.; Strassmann, K.; Schnur, R.; Schnitzler, K.-G.; Roeckner, E.; Reick, C.; Rayner, P.; Raddatz, T.; Matthews, H. D.; Lindsay, K.; Knorr, W.; Kawamiya, M.; Kato, T.; Joos, F.; Jones, C.; John, J.; Bala, G.; Fung, I.; Eby, M.; Doney, S.; Cadule, P.; Brovkin, V.; von Bloh, W.; Bopp, L.; Betts, R.; Cox, P.; Friedlingstein, P. (2006). "Climate–Carbon Cycle Feedback Analysis: Results from the C4MIP Model Intercomparison". Journal of Climate. 19 (14): 3337–3353. Bibcode:2006JCli...19.3337F. doi: 10.1175/JCLI3800.1 . hdl: 10036/68733 .
  16. Kuzyakov, Yakov; Horwath, William R.; Dorodnikov, Maxim; Blagodatskaya, Evgenia (2019). "Review and synthesis of the effects of elevated atmospheric CO2 on soil processes: No changes in pools, but increased fluxes and accelerated cycles". Soil Biology and Biochemistry. 128: 66–78. doi:10.1016/j.soilbio.2018.10.005.
  17. 1 2 Liang, Chao; Schimel, Joshua P.; Jastrow, Julie D. (2017). "The importance of anabolism in microbial control over soil carbon storage". Nature Microbiology. 2 (8): 17105. doi:10.1038/nmicrobiol.2017.105. PMID   28741607. S2CID   9992380.
  18. Bonkowski, Michael (2004). "Protozoa and plant growth: The microbial loop in soil revisited". New Phytologist. 162 (3): 617–631. doi: 10.1111/j.1469-8137.2004.01066.x . PMID   33873756.
  19. Sulman, Benjamin N.; Phillips, Richard P.; Oishi, A. Christopher; Shevliakova, Elena; Pacala, Stephen W. (2014). "Microbe-driven turnover offsets mineral-mediated storage of soil carbon under elevated CO2". Nature Climate Change. 4 (12): 1099–1102. Bibcode:2014NatCC...4.1099S. doi:10.1038/nclimate2436.
  20. Harmon, Mark E.; Silver, Whendee L.; Fasth, Becky; Chen, HUA; Burke, Ingrid C.; Parton, William J.; Hart, Stephen C.; Currie, William S. (2009). "Long-term patterns of mass loss during the decomposition of leaf and fine root litter: An intersite comparison". Global Change Biology. 15 (5): 1320–1338. Bibcode:2009GCBio..15.1320H. doi:10.1111/j.1365-2486.2008.01837.x. hdl: 2027.42/74496 .
  21. Qualls, Robert (2016). "Long-Term (13 Years) Decomposition Rates of Forest Floor Organic Matter on Paired Coniferous and Deciduous Watersheds with Contrasting Temperature Regimes". Forests. 7 (12): 231. doi: 10.3390/f7100231 .
  22. Lange, Markus; Eisenhauer, Nico; Sierra, Carlos A.; Bessler, Holger; Engels, Christoph; Griffiths, Robert I.; Mellado-Vázquez, Perla G.; Malik, Ashish A.; Roy, Jacques; Scheu, Stefan; Steinbeiss, Sibylle; Thomson, Bruce C.; Trumbore, Susan E.; Gleixner, Gerd (2015). "Plant diversity increases soil microbial activity and soil carbon storage" (PDF). Nature Communications. 6: 6707. Bibcode:2015NatCo...6.6707L. doi:10.1038/ncomms7707. PMID   25848862.
  23. Stewart, Catherine E.; Paustian, Keith; Conant, Richard T.; Plante, Alain F.; Six, Johan (2007). "Soil carbon saturation: Concept, evidence and evaluation". Biogeochemistry. 86: 19–31. doi:10.1007/s10533-007-9140-0. S2CID   97153551.

Bibliography