Dissolved organic carbon

Last updated

Net ocean DOC production and export fluxes
Net DOC production (NDP) in the upper 74 metres (a) and net DOC export (NDX) below 74 metres (b). At steady state, the global summation of NDX is equal to that of NDP, and is 2.31 +- 0.60 PgC yr. Ocean DOC production and export fluxes.png
Net ocean DOC production and export fluxes
Net DOC production (NDP) in the upper 74 metres (a) and net DOC export (NDX) below 74 metres (b). At steady state, the global summation of NDX is equal to that of NDP, and is 2.31 ± 0.60 PgC yr.

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. [2] The fraction remaining on the filter is called particulate organic carbon (POC). [3]

Contents

Dissolved organic matter (DOM) is a closely related term often used interchangeably with DOC. While DOC refers specifically to the mass of carbon in the dissolved organic material, DOM refers to the total mass of the dissolved organic matter. So DOM also includes the mass of other elements present in the organic material, such as nitrogen, oxygen and hydrogen. DOC is a component of DOM and there is typically about twice as much DOM as DOC. [4] Many statements that can be made about DOC apply equally to DOM, and vice versa.

DOC is abundant in marine and freshwater systems and is one of the greatest cycled reservoirs of organic matter on Earth, accounting for the same amount of carbon as in the atmosphere and up to 20% of all organic carbon. [5] In general, organic carbon compounds are the result of decomposition processes from dead organic matter including plants and animals. DOC can originate from within or outside any given body of water. DOC originating from within the body of water is known as autochthonous DOC and typically comes from aquatic plants or algae, while DOC originating outside the body of water is known as allochthonous DOC and typically comes from soils or terrestrial plants. [6] When water originates from land areas with a high proportion of organic soils, these components can drain into rivers and lakes as DOC.

The marine DOC pool is important for the functioning of marine ecosystems because they are at the interface between the chemical and the biological worlds. DOC fuels marine food webs, and is a major component of the Earth's carbon cycling. [7]

Overview

Size and classification of marine particles
Adapted from Simon et al., 2002. Size and classification of marine particles.png
Size and classification of marine particles
Adapted from Simon et al., 2002.
Colour differences in DOC collected from coastal waters
Filtered (0.2 mm) coastal marine waters collected at various locations around the United Kingdom. The differences in colour is due to the range of soil-derived carbon input to the coastal water, with dark brown (left) indicating a high soil-derived carbon contribution and near-clear water (right) indicating a low soil-derived carbon contribution. Colour differences in filtered coastal waters.jpg
Colour differences in DOC collected from coastal waters
Filtered (0.2 μm) coastal marine waters collected at various locations around the United Kingdom. The differences in colour is due to the range of soil-derived carbon input to the coastal water, with dark brown (left) indicating a high soil-derived carbon contribution and near-clear water (right) indicating a low soil-derived carbon contribution.

DOC is a basic nutrient, supporting growth of microorganisms and plays an important role in the global carbon cycle through the microbial loop. [10] In some organisms (stages) that do not feed in the traditional sense, dissolved matter may be the only external food source. [11] Moreover, DOC is an indicator of organic loadings in streams, as well as supporting terrestrial processing (e.g., within soil, forests, and wetlands) of organic matter. Dissolved organic carbon has a high proportion of biodegradable dissolved organic carbon (BDOC) in first order streams compared to higher order streams. In the absence of extensive wetlands, bogs, or swamps, baseflow concentrations of DOC in undisturbed watersheds generally range from approximately 1 to 20 mg/L carbon. [12] Carbon concentrations considerably vary across ecosystems. For example, the Everglades may be near the top of the range and the middle of oceans may be near the bottom. Occasionally, high concentrations of organic carbon indicate anthropogenic influences, but most DOC originates naturally. [13]

The BDOC fraction consists of organic molecules that heterotrophic bacteria can use as a source of energy and carbon. [14] Some subset of DOC constitutes the precursors of disinfection byproducts for drinking water. [15] BDOC can contribute to undesirable biological regrowth within water distribution systems. [16]

The dissolved fraction of total organic carbon (TOC) is an operational classification. Many researchers use the term "dissolved" for compounds that pass through a 0.45 μm filter, but 0.22 μm filters have also been used to remove higher colloidal concentrations. [2]

A practical definition of dissolved typically used in marine chemistry is all substances that pass through a GF/F filter, which has a nominal pore size of approximately 0.7 μm (Whatman glass microfiber filter, 0.6–0.8 μm particle retention [17] ). The recommended procedure is the HTCO technique, which calls for filtration through pre-combusted glass fiber filters, typically the GF/F classification. [18]

Labile and recalcitrant

Dissolved organic matter can be classified as labile or as recalcitrant, depending on its reactivity. Recalcitrant DOC is also called refractory DOC, and these terms seem to be used interchangeably in the context of DOC. Depending on the origin and composition of DOC, its behavior and cycling are different; the labile fraction of DOC decomposes rapidly through microbially or photochemically mediated processes, whereas refractory DOC is resistant to degradation and can persist in the ocean for millennia. In the coastal ocean, organic matter from terrestrial plant litter or soils appears to be more refractory [19] and thus often behaves conservatively. In addition, refractory DOC is produced in the ocean by the bacterial transformation of labile DOC, which reshapes its composition. [20] [21] [22]

Due to the continuous production and degradation in natural systems, the DOC pool contains a spectrum of reactive compounds each with their own reactivity, [23] that have been divided into fractions from labile to recalcitrant, depending on the turnover times, [24] as shown in the following table...

DOC pool spectrum from labile to recalcitrant [23] [24]
DOC fractionacronymturnover timeamount
labileDOCLhours to days< 200 Tg C
semi-labileDOCSLweeks to months~600 Tg C
semi-recalcitrantDOCSRdecades~1400 Tg C
recalcitrantDOCRthousands of years~63000 Tg C
highly resistanttens of thousands of years

This wide range in turnover or degradation times has been linked with the chemical composition, structure and molecular size, [25] [26] but degradation also depends on the environmental conditions (e.g., nutrients), prokaryote diversity, redox state, iron availability, mineral-particle associations, temperature, sun-light exposure, biological production of recalcitrant compounds, and the effect of priming or dilution of individual molecules. [25] [27] [28] [29] [30] [31] For example, lignin can be degraded in aerobic soils but is relatively recalcitrant in anoxic marine sediments. [32] This example shows bioavailability varies as a function of the ecosystem's properties. Accordingly, even normally ancient and recalcitrant compounds, such as petroleum, carboxyl-rich alicyclic molecules, can be degraded in the appropriate environmental setting. [33] [34]

Terrestrial ecosystems

Soil

Soil DOC sources and sinks [35]
DOM inputs, processes and losses in the soil system.jpg
Sources and sinks of dissolved organic carbon in the soil system
Groundwater DOC sources
Origins and bioavailability of dissolved organic matter in groundwater.webp
Origins and bioavailability of DOC in groundwater [36]
DOM: dissolved organic matter
Freshwater DOC sources and sinks [37]
Carbon forms and interactions in freshwater ecosystems.jpg
DOC and POC — DIC and PIC
Inland waters primarily receive carbon from terrestrial ecosystems. [38] This carbon (1.9 Pg C y−1) is transported to the oceans (0.9 Pg C y−1), buried in the sediments (0.2 Pg C y−1) or emitted as CO2 (0.8 Pg C y−1). [39] More recent estimations are different: In 2013, Raymond et al. claimed CO2 emission from inland waters can be as high as 2.1 Pg C y−1. [40]
                   P = photosynthesis                    R = respiration

Dissolved organic matter (DOM) is one of the most active and mobile carbon pools and has an important role in global carbon cycling. [41] In addition, dissolved organic carbon (DOC) affects the soil negative electrical charges denitrification process, acid-base reactions in the soil solution, retention and translocation of nutrients (cations), and immobilization of heavy metals and xenobiotics. [42] Soil DOM can be derived from different sources (inputs), such as atmospheric carbon dissolved in rainfall, litter and crop residues, manure, root exudates, and decomposition of soil organic matter (SOM). In the soil, DOM availability depends on its interactions with mineral components (e.g., clays, Fe and Al oxides) modulated by adsorption and desorption processes. [43] It also depends on SOM fractions (e.g., stabilized organic molecules and microbial biomass) by mineralization and immobilization processes. In addition, the intensity of these interactions changes according to soil inherent properties, [44] land use, and crop management. [43] [35]

During the decomposition of organic material, most carbon is lost as CO2 to the atmosphere by microbial oxidation. Soil type and landscape slope, leaching, and runoff are also important processes associated to DOM losses in the soil. [45] In well-drained soils, leached DOC can reach the water table and release nutrients and pollutants that can contaminate groundwater, [45] [46] whereas runoff transports DOM and xenobiotics to other areas, rivers, and lakes. [35]

Groundwater

Precipitation and surface water leaches dissolved organic carbon (DOC) from vegetation and plant litter and percolates through the soil column to the saturated zone. The concentration, composition, and bioavailability of DOC are altered during transport through the soil column by various physicochemical and biological processes, including sorption, desorption, biodegradation and biosynthesis. Hydrophobic molecules are preferentially partitioned onto soil minerals and have a longer retention time in soils than hydrophilic molecules. The hydrophobicity and retention time of colloids and dissolved molecules in soils are controlled by their size, polarity, charge, and bioavailability. Bioavailable DOM is subjected to microbial decomposition, resulting in a reduction in size and molecular weight. Novel molecules are synthesized by soil microbes, and some of these metabolites enter the DOC reservoir in groundwater. [36]

Freshwater ecosystems

Aquatic carbon occurs in different forms. Firstly, a division is made between organic and inorganic carbon. Organic carbon is a mixture of organic compounds originating from detritus or primary producers. It can be divided into POC (particulate organic carbon; particles > 0.45 μm) and DOC (dissolved organic carbon; particles < 0.45 μm). DOC usually makes up 90% of the total amount of aquatic organic carbon. Its concentration ranges from 0.1 to >300 mg L−1. [47]

Likewise, inorganic carbon also consists of a particulate (PIC) and a dissolved phase (DIC). PIC mainly consists of carbonates (e.g., CaCO3), DIC consists of carbonate (CO32-), bicarbonate (HCO3), CO2 and a negligibly small fraction of carbonic acid (H2CO3). The inorganic carbon compounds exist in equilibrium that depends on the pH of the water. [48] DIC concentrations in freshwater range from about zero in acidic waters to 60 mg C L−1 in areas with carbonate-rich sediments. [49]

POC can be degraded to form DOC; DOC can become POC by flocculation. Inorganic and organic carbon are linked through aquatic organisms. CO2 is used in photosynthesis (P) by for instance macrophytes, produced by respiration (R), and exchanged with the atmosphere. Organic carbon is produced by organisms and is released during and after their life; e.g., in rivers, 1–20% of the total amount of DOC is produced by macrophytes. [38] Carbon can enter the system from the catchment and is transported to the oceans by rivers and streams. There is also exchange with carbon in the sediments, e.g., burial of organic carbon, which is important for carbon sequestration in aquatic habitats. [50]

Aquatic systems are very important in global carbon sequestration; e.g., when different European ecosystems are compared, inland aquatic systems form the second largest carbon sink (19–41 Tg C y−1); only forests take up more carbon (125–223 Tg C y−1). [51] [37]

Marine ecosystems

Ocean DOC sources and sinks [7]
Main sources of ocean dissolved organic carbon.jpg
Simplified view of the main sources (black text; underlined are the allochthonous sources) and sinks (yellow text) of the oceanic dissolved organic carbon (DOC) pool.
Main sources
Most commonly referred sources of DOC are: atmospheric (e.g., rain and dust), terrestrial (e.g., rivers), primary producers (e.g., microalgae, cyanobacteria, macrophytes), groundwater, food chain processes (e.g., zooplankton grazing), and benthic fluxes (exchange of DOC across the sediment-water interface but also from hydrothermal vents). [7]
Main sinks
The four main processes removing DOC from the water column are: photodegradation (particularly UV-radiation – though sometimes photodegradation "transforms" DOC rather than removing it, ending up with higher molecular weight complex molecules), microbial (mainly by prokaryotes), aggregation (primarily when river and seawater mixes) and thermal degradation (in e.g., hydrothermal systems). [7]

Sources

In marine systems DOC originates from either autochthonous or allochthonous sources. Autochthonous DOC is produced within the system, primarily by plankton organisms [52] [53] and in coastal waters additionally by benthic microalgae, benthic fluxes, and macrophytes, [54] whereas allochthonous DOC is mainly of terrestrial origin supplemented by groundwater and atmospheric inputs. [55] [56] In addition to soil derived humic substances, terrestrial DOC also includes material leached from plants exported during rain events, emissions of plant materials to the atmosphere and deposition in aquatic environments (e.g., volatile organic carbon and pollens), and also thousands of synthetic human-made organic chemicals that can be measured in the ocean at trace concentrations. [57] [56] [7]

Dissolved organic carbon (DOC) represents one of the Earth's major carbon pools. It contains a similar amount of carbon as the atmosphere and exceeds the amount of carbon bound in marine biomass by more than two-hundred times. [58] DOC is mainly produced in the near-surface layers during primary production and zooplankton grazing processes. [59] Other sources of marine DOC are dissolution from particles, [59] terrestrial and hydrothermal vent input, [60] and microbial production. Prokaryotes (bacteria and archaea) contribute to the DOC pool via release of capsular material, exopolymers, and hydrolytic enzymes, [59] as well as via mortality (e.g. viral shunt). Prokaryotes are also the main decomposers of DOC, although for some of the most recalcitrant forms of DOC very slow abiotic degradation in hydrothermal systems [59] or possibly sorption to sinking particles [24] may be the main removal mechanism. Mechanistic knowledge about DOC-microbe-interactions is crucial to understand the cycling and distribution of this active carbon reservoir. [61]

Phytoplankton

Phytoplankton produces DOC by extracellular release commonly accounting between 5 and 30% of their total primary production, [62] although this varies from species to species. [63] Nonetheless, this release of extracellular DOC is enhanced under high light and low nutrient levels, and thus should increase relatively from eutrophic to oligotrophic areas, probably as a mechanism for dissipating cellular energy. [64] Phytoplankton can also produce DOC by autolysis during physiological stress situations e.g., nutrient limitation. [65] Other studies have demonstrated DOC production in association with meso- and macro-zooplankton feeding on phytoplankton and bacteria. [66] [7]

Zooplankton

Zooplankton-mediated release of DOC occurs through sloppy feeding, excretion and defecation which can be important energy sources for microbes. [67] [66] Such DOC production is largest during periods with high food concentration and dominance of large zooplankton species. [68] [7]

Bacteria and viruses

Bacteria are often viewed as the main consumers of DOC, but they can also produce DOC during cell division and viral lysis. [69] [70] [52] The biochemical components of bacteria are largely the same as other organisms, but some compounds from the cell wall are unique and are used to trace bacterial derived DOC (e.g., peptidoglycan). These compounds are widely distributed in the ocean, suggesting that bacterial DOC production could be important in marine systems. [71] Viruses are the most abundant life forms in the oceans infecting all life forms including algae, bacteria and zooplankton. [72] After infection, the virus either enters a dormant (lysogenic) or productive (lytic) state. [73] The lytic cycle causes disruption of the cell(s) and release of DOC. [74] [7]

DOC net production, transport and export in the ocean
Regions of significant net DOC production (broad arrows) include coastal and equatorial upwelling regions that support much of the global new production. DOC is transported into and around the subtropical gyres with the wind-driven surface circulation. Export takes place if exportable DOC (elevated concentrations indicated by dark blue fields) is present during overturning of the water column. precursor for deep and intermediate water mass formation. DOC is also exported with subduction in the gyres. In regions where DOCenriched subtropical water is prevented by polar frontal systems from serving as a precursor for overturning circulation (such as at the sites of Antarctic Bottom Water formation in the Southern Ocean) DOC export is a weak component of the biological pump. Waters south of the Antarctic Polar Front lack significant exportable DOC (depicted by light blue field) during winter. DOC net production, transport and export in the ocean.png
DOC net production, transport and export in the ocean
Regions of significant net DOC production (broad arrows) include coastal and equatorial upwelling regions that support much of the global new production. DOC is transported into and around the subtropical gyres with the wind-driven surface circulation. Export takes place if exportable DOC (elevated concentrations indicated by dark blue fields) is present during overturning of the water column. precursor for deep and intermediate water mass formation. DOC is also exported with subduction in the gyres. In regions where DOCenriched subtropical water is prevented by polar frontal systems from serving as a precursor for overturning circulation (such as at the sites of Antarctic Bottom Water formation in the Southern Ocean) DOC export is a weak component of the biological pump. Waters south of the Antarctic Polar Front lack significant exportable DOC (depicted by light blue field) during winter.
Simplified microbial food web in the sunlit ocean
Left side: classic description of the carbon flow from photosynthetic algae to grazers and higher trophic levels in the food chain.
Right side: microbial loop, with bacteria using dissolved organic carbon to gain biomass, which then re-enters the classic carbon flow through protists. Simplified microbial food web in the sunlit ocean.png
Simplified microbial food web in the sunlit ocean
Left side: classic description of the carbon flow from photosynthetic algae to grazers and higher trophic levels in the food chain.
Right side: microbial loop, with bacteria using dissolved organic carbon to gain biomass, which then re-enters the classic carbon flow through protists.
Dissolved organic carbon (DOC) fluxes in the surface, mesopelagic, and interior ocean
In panel (A) oceanic DOC stocks are shown in black circles with red font and units are Pg-C. DOC fluxes are shown in black and white font and units are either Tg-C yr or Pg-C yr . Letters in arrows and associated flux values correspond to descriptions displayed in (B), which lists sources and sinks of oceanic DOC. Dissolved organic carbon fluxes in the surface, mesopelagic, and interior ocean.jpg
Dissolved organic carbon (DOC) fluxes in the surface, mesopelagic, and interior ocean
In panel (A) oceanic DOC stocks are shown in black circles with red font and units are Pg-C. DOC fluxes are shown in black and white font and units are either Tg-C yr or Pg-C yr . Letters in arrows and associated flux values correspond to descriptions displayed in (B), which lists sources and sinks of oceanic DOC.

Macrophytes

Marine macrophytes (i.e., macroalgae and seagrass) are highly productive and extend over large areas in coastal waters but their production of DOC has not received much attention. Macrophytes release DOC during growth with a conservative estimate (excluding release from decaying tissues) suggesting that macroalgae release between 1-39% of their gross primary production, [79] [80] while seagrasses release less than 5% as DOC of their gross primary production. [81] The released DOC has been shown to be rich in carbohydrates, with rates depending on temperature and light availability. [82] Globally the macrophyte communities have been suggested to produce ~160 Tg C yr−1 of DOC, which is approximately half the annual global river DOC input (250 Tg C yr−1). [82] [7]

Marine sediments

Peatland river water draining into coastal waters
South-East Asia is home to one of the world's largest stores of tropical peatland and accounts for roughly 10 % of the global land-to-sea dissolved organic carbon (DOC) flux. The rivers carry high coloured dissolved organic matter (CDOM) concentrations, shown here interfacing with ocean shelf water. Interface between peatland draining river water and coastal waters.png
Peatland river water draining into coastal waters
South-East Asia is home to one of the world's largest stores of tropical peatland and accounts for roughly 10 % of the global land-to-sea dissolved organic carbon (DOC) flux. The rivers carry high coloured dissolved organic matter (CDOM) concentrations, shown here interfacing with ocean shelf water.

Marine sediments represent the main sites of OM degradation and burial in the ocean, hosting microbes in densities up to 1000 times higher than found in the water column. [84] The DOC concentrations in sediments are often an order of magnitude higher than in the overlying water column. [85] This concentration difference results in a continued diffusive flux and suggests that sediments are a major DOC source releasing 350 Tg C yr−1, which is comparable to the input of DOC from rivers. [86] This estimate is based on calculated diffusive fluxes and does not include resuspension events which also releases DOC [87] and therefore the estimate could be conservative. Also, some studies have shown that geothermal systems and petroleum seepage contribute with pre-aged DOC to the deep ocean basins, [88] [89] but consistent global estimates of the overall input are currently lacking. Globally, groundwaters account for an unknown part of the freshwater DOC flux to the oceans. [90] The DOC in groundwater is a mixture of terrestrial, infiltrated marine, and in situ microbially produced material. [91] This flux of DOC to coastal waters could be important, as concentrations in groundwater are generally higher than in coastal seawater, [92] but reliable global estimates are also currently lacking. [7]

Sinks

The main processes that remove DOC from the ocean water column are: (1) Thermal degradation in e.g., submarine hydrothermal systems; [93] (2) bubble coagulation and abiotic flocculation into microparticles [94] or sorption to particles; [95] (3) abiotic degradation via photochemical reactions; [96] [97] and (4) biotic degradation by heterotrophic marine prokaryotes. [98] It has been suggested that the combined effects of photochemical and microbial degradation represent the major sinks of DOC. [99] [7]

Thermal degradation

Thermal degradation of DOC has been found at high-temperature hydrothermal ridge-flanks, where outflow DOC concentrations are lower than in the inflow. While the global impact of these processes has not been investigated, current data suggest it is a minor DOC sink. [93] Abiotic DOC flocculation is often observed during rapid (minutes) shifts in salinity when fresh and marine waters mix. [101] Flocculation changes the DOC chemical composition, by removing humic compounds and reducing molecular size, transforming DOC to particulate organic flocs which can sediment and/or be consumed by grazers and filter feeders, but it also stimulates the bacterial degradation of the flocculated DOC. [102] The impacts of flocculation on the removal of DOC from coastal waters are highly variable with some studies suggesting it can remove up to 30% of the DOC pool, [103] [104] while others find much lower values (3–6%; [105] ). Such differences could be explained by seasonal and system differences in the DOC chemical composition, pH, metallic cation concentration, microbial reactivity, and ionic strength. [101] [106] [7]

CDOM

The colored fraction of DOC (CDOM) absorbs light in the blue and UV-light range and therefore influences plankton productivity both negatively by absorbing light, that otherwise would be available for photosynthesis, and positively by protecting plankton organisms from harmful UV-light. [107] [108] However, as the impact of UV damage and ability to repair is extremely variable, there is no consensus on how UV-light changes might impact overall plankton communities. [109] [110] The CDOM absorption of light initiates a complex range of photochemical processes, which can impact nutrient, trace metal and DOC chemical composition, and promote DOC degradation. [97]

Photodegradation

Photodegradation involves the transformation of CDOM into smaller and less colored molecules (e.g., organic acids), or into inorganic carbon (CO, CO2), and nutrient salts (NH4, HPO2−
4
). [111] [96] [112] Therefore, it generally means that photodegradation transforms recalcitrant into labile DOC molecules that can be rapidly used by prokaryotes for biomass production and respiration. However, it can also increase CDOM through the transformation of compounds such as triglycerides, into more complex aromatic compounds, [113] [114] which are less degradable by microbes. Moreover, UV radiation can produce e.g., reactive oxygen species, which are harmful to microbes. [115] The impact of photochemical processes on the DOC pool depends also on the chemical composition, [116] with some studies suggesting that recently produced autochthonous DOC becomes less bioavailable while allochthonous DOC becomes more bioavailable to prokaryotes after sunlight exposure, albeit others have found the contrary. [117] [118] [119] Photochemical reactions are particularly important in coastal waters which receive high loads of terrestrial derived CDOM, with an estimated ~20–30% of terrestrial DOC being rapidly photodegraded and consumed. [120] Global estimates also suggests that in marine systems photodegradation of DOC produces ~180 Tg C yr−1 of inorganic carbon, with an additional 100 Tg C yr−1 of DOC made more available to microbial degradation. [96] [121] Another attempt at global ocean estimates also suggest that photodegradation (210 Tg C yr−1) is approximately the same as the annual global input of riverine DOC (250 Tg C yr−1; [122] ), while others suggest that direct photodegradation exceeds the riverine DOC inputs. [123] [124] [7]

Recalcitrant DOC

Change in the composition of DOC with depth Vertical Dissolved Organic Carbon Distribution.PNG
Change in the composition of DOC with depth

DOC is conceptually divided into labile DOC, which is rapidly taken up by heterotrophic microbes, and the recalcitrant DOC reservoir, which has accumulated in the ocean (following a definition by Hansell). [24] As a consequence of its recalcitrance, the accumulated DOC reaches average radiocarbon ages between 1,000 and 4,000 years in surface waters, and between 3,000 and 6,000 years in the deep ocean, [125] indicating that it persists through several deep ocean mixing cycles between 300 and 1,400 years each. [126] Behind these average radiocarbon ages, a large spectrum of ages is hidden. Follett et al. showed DOC comprises a fraction of modern radiocarbon age, as well as DOC reaching radiocarbon ages of up to 12,000 years. [127] [61]

Distribution

More precise measurement techniques developed in the late 1990s have allowed for a good understanding of how dissolved organic carbon is distributed in marine environments both vertically and across the surface. [128] It is now understood that dissolved organic carbon in the ocean spans a range from very labile to very recalcitrant (refractory). The labile dissolved organic carbon is mainly produced by marine organisms and is consumed in the surface ocean, and consists of sugars, proteins, and other compounds that are easily used by marine bacteria. [129] Recalcitrant dissolved organic carbon is evenly spread throughout the water column and consists of high molecular weight and structurally complex compounds that are difficult for marine organisms to use such as the lignin, pollen, or humic acids. As a result, the observed vertical distribution consists of high concentrations of labile DOC in the upper water column and low concentrations at depth. [130]

In addition to vertical distributions, horizontal distributions have been modeled and sampled as well. [131] In the surface ocean at a depth of 30 meters, the higher dissolved organic carbon concentrations are found in the South Pacific Gyre, the South Atlantic Gyre, and the Indian Ocean. At a depth of 3,000 meters, highest concentrations are in the North Atlantic Deep Water where dissolved organic carbon from the high concentration surface ocean is removed to depth. While in the northern Indian Ocean high DOC is observed due to high fresh water flux and sediments. Since the time scales of horizontal motion along the ocean bottom are in the thousands of years, the refractory dissolved organic carbon is slowly consumed on its way from the North Atlantic and reaches a minimum in the North Pacific. [131]

As emergent

Dissolved organic matter is a heterogeneous pool of thousands, likely millions, of organic compounds. These compounds differ not only in composition and concentration (from pM to μM), but also originate from various organisms (phytoplankton, zooplankton, and bacteria) and environments (terrestrial vegetation and soils, coastal fringe ecosystems) and may have been produced recently or thousands of years ago. Moreover, even organic compounds deriving from the same source and of the same age may have been subjected to different processing histories prior to accumulating within the same pool of DOM. [78]

Interior ocean DOM is a highly modified fraction that remains after years of exposure to sunlight, utilization by heterotrophs, flocculation and coagulation, and interaction with particles. Many of these processes within the DOM pool are compound- or class-specific. For example, condensed aromatic compounds are highly photosensitive, [132] whereas proteins, carbohydrates, and their monomers are readily taken up by bacteria. [133] [134] [135] Microbes and other consumers are selective in the type of DOM they utilize and typically prefer certain organic compounds over others. Consequently, DOM becomes less reactive as it is continually reworked. Said another way, the DOM pool becomes less labile and more refractory with degradation. As it is reworked, organic compounds are continually being added to the bulk DOM pool by physical mixing, exchange with particles, and/or production of organic molecules by the consumer community. [70] [21] [136] [100] As such, the compositional changes that occur during degradation are more complex than the simple removal of more labile components and resultant accumulation of remaining, less labile compounds. [78]

Dissolved organic matter recalcitrance (i.e., its overall reactivity toward degradation and/or utilization) is therefore an emergent property. The perception of DOM recalcitrance changes during organic matter degradation and in conjunction with any other process that removes or adds organic compounds to the DOM pool under consideration. [78]

The surprising resistance of high concentrations of DOC to microbial degradation has been addressed by several hypotheses. [137] The prevalent notion is that the recalcitrant fraction of DOC has certain chemical properties, which prevent decomposition by microbes ("intrinsic stability hypothesis"). An alternative or additional explanation is given by the "dilution hypothesis", that all compounds are labile, but exist in concentrations individually too low to sustain microbial populations but collectively form a large pool. [138] The dilution hypothesis has found support in recent experimental and theoretical studies. [139] [140] [61]

DOM isolation and analysis

DOM is found in low concentrations in nature for direct analysis with NMR or MS. Moreover, DOM samples often contain high concentrations of inorganic salts that are incompatible with such techniques. [141] Therefore, it is necessary a concentration and isolation step of the sample. [141] [142] The most used isolation techniques are ultrafiltration, reverse osmosis, and solid-phase extraction. [143] Among them solid-phase extraction is considered as the cheapest and easiest technique. [142]

See also

Related Research Articles

<span class="mw-page-title-main">Humus</span> Organic matter in soils resulting from decay of plant and animal materials

In classical soil science, humus is the dark organic matter in soil that is formed by the decomposition of plant and animal matter. It is a kind of soil organic matter. It is rich in nutrients and retains moisture in the soil. Humus is the Latin word for "earth" or "ground".

<span class="mw-page-title-main">Carbon cycle</span> Natural processes of carbon exchange

The carbon cycle is that part of the biogeochemical cycle by which carbon is exchanged among the biosphere, pedosphere, geosphere, hydrosphere, and atmosphere of Earth. Other major biogeochemical cycles include the nitrogen cycle and the water cycle. Carbon is the main component of biological compounds as well as a major component of many minerals such as limestone. The carbon cycle comprises a sequence of events that are key to making Earth capable of sustaining life. It describes the movement of carbon as it is recycled and reused throughout the biosphere, as well as long-term processes of carbon sequestration (storage) to and release from carbon sinks.

<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).

The bathypelagic zone or bathyal zone is the part of the open ocean that extends from a depth of 1,000 to 4,000 m below the ocean surface. It lies between the mesopelagic above and the abyssopelagic below. The bathypelagic is also known as the midnight zone because of the lack of sunlight; this feature does not allow for photosynthesis-driven primary production, preventing growth of phytoplankton or aquatic plants. Although larger by volume than the photic zone, human knowledge of the bathypelagic zone remains limited by ability to explore the deep ocean.

Organic matter, organic material, or natural organic matter refers to the large source of carbon-based compounds found within natural and engineered, terrestrial, and aquatic environments. It is matter composed of organic compounds that have come from the feces and remains of organisms such as plants and animals. Organic molecules can also be made by chemical reactions that do not involve life. Basic structures are created from cellulose, tannin, cutin, and lignin, along with other various proteins, lipids, and carbohydrates. Organic matter is very important in the movement of nutrients in the environment and plays a role in water retention on the surface of the planet.

<span class="mw-page-title-main">Dissolved inorganic carbon</span> Sum of inorganic carbon species in a solution

Dissolved inorganic carbon (DIC) is the sum of the aqueous species of inorganic carbon in a solution. Carbon compounds can be distinguished as either organic or inorganic, and as dissolved or particulate, depending on their composition. Organic carbon forms the backbone of key component of organic compounds such as – proteins, lipids, carbohydrates, and nucleic acids.

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.

f-ratio (oceanography) In oceanic biogeochemistry, the fraction of total primary production fuelled by nitrate

In oceanic biogeochemistry, the f-ratio is the fraction of total primary production fuelled by nitrate. The ratio was originally defined by Richard Eppley and Bruce Peterson in one of the first papers estimating global oceanic production. This fraction was originally believed significant because it appeared to directly relate to the sinking (export) flux of organic marine snow from the surface ocean by the biological pump. However, this interpretation relied on the assumption of a strong depth-partitioning of a parallel process, nitrification, that more recent measurements has questioned.

Heterotrophic picoplankton is the fraction of plankton composed by cells between 0.2 and 2 μm that do not perform photosynthesis. They form an important component of many biogeochemical cycles.

<span class="mw-page-title-main">Carbon-to-nitrogen ratio</span>

A carbon-to-nitrogen ratio is a ratio of the mass of carbon to the mass of nitrogen in organic residues. It can, amongst other things, be used in analysing sediments and soil including soil organic matter and soil amendments such as compost.

<span class="mw-page-title-main">Microbial loop</span> Trophic pathway in marine microbial ecosystems

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. in 1983 to include the role played by bacteria in the carbon and nutrient cycles of the marine environment.

<span class="mw-page-title-main">Colored dissolved organic matter</span> Optically measurable component of the dissolved organic matter in water

Colored dissolved organic matter (CDOM) is the optically measurable component of dissolved organic matter in water. Also known as chromophoric dissolved organic matter, yellow substance, and gelbstoff, CDOM occurs naturally in aquatic environments and is a complex mixture of many hundreds to thousands of individual, unique organic matter molecules, which are primarily leached from decaying detritus and organic matter. CDOM most strongly absorbs short wavelength light ranging from blue to ultraviolet, whereas pure water absorbs longer wavelength red light. Therefore, water with little or no CDOM, such as the open ocean, appears blue. Waters containing high amounts of CDOM can range from brown, as in many rivers, to yellow and yellow-brown in coastal waters. In general, CDOM concentrations are much higher in fresh waters and estuaries than in the open ocean, though concentrations are highly variable, as is the estimated contribution of CDOM to the total dissolved organic matter pool.

<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">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">Oceanic carbon cycle</span> Ocean/atmosphere carbon exchange process

The oceanic carbon cycle is composed of processes that exchange carbon between various pools within the ocean as well as between the atmosphere, Earth interior, and the seafloor. The carbon cycle is a result of many interacting forces across multiple time and space scales that circulates carbon around the planet, ensuring that carbon is available globally. The Oceanic carbon cycle is a central process to the global carbon cycle and contains both inorganic carbon and organic carbon. Part of the marine carbon cycle transforms carbon between non-living and living matter.

<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">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">Martin curve</span> Mathematical representation of particulate organic carbon export to ocean floor

The Martin curve is a power law used by oceanographers to describe the export to the ocean floor of particulate organic carbon (POC). The curve is controlled with two parameters: the reference depth in the water column, and a remineralisation parameter which is a measure of the rate at which the vertical flux of POC attenuates. It is named after the American oceanographer John Martin.

Elizabeth A. Canuel is a chemical oceanographer known for her work on organic carbon cycling in aquatic environments. She is the Chancellor Professor of Marine Science at the College of William & Mary and is an elected fellow of the Geochemical Society and the European Association of Geochemistry.

References

  1. Roshan, S. and DeVries, T. (2017) "Efficient dissolved organic carbon production and export in the oligotrophic ocean". Nature Communications, 8(1): 1–8. doi : 10.1038/s41467-017-02227-3.
  2. 1 2 "Organic Carbon". Bio-geochemical Methods. Retrieved 27 November 2018.
  3. Kenny, Jonathan E.; Bida, Morgan; Pagano, Todd (October 2014). "Trends in Levels of Allochthonous Dissolved Organic Carbon in Natural Water: A Review of Potential Mechanisms under a Changing Climate". Water. 6 (10): 2862–2897. doi: 10.3390/w6102862 .
  4. Moody, C.S. and Worrall, F. (2017) "Modeling rates of DOC degradation using DOM composition and hydroclimatic variables". Journal of Geophysical Research: Biogeosciences, 122(5): 1175–1191. doi : 10.1002/2016JG003493.
  5. Hedges, John I. (3 December 1991). "Global biogeochemical cycles: progress and problems" (PDF). Marine Chemistry. 39 (1–3): 67–93. doi:10.1016/0304-4203(92)90096-s.
  6. Kritzberg, Emma S.; Cole, Jonathan J.; Pace, Michael L.; Granéli, Wilhelm; Bade, Darren L. (March 2004). "Autochthonous versus allochthonous carbon sources of bacteria: Results from whole-lake 13C addition experiments" (PDF). Limnology and Oceanography. 49 (2): 588–596. Bibcode:2004LimOc..49..588K. doi:10.4319/lo.2004.49.2.0588. ISSN   0024-3590. S2CID   15021562.
  7. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Lønborg, C., Carreira, C., Jickells, T. and Álvarez-Salgado, X.A. (2020) "Impacts of global change on ocean dissolved organic carbon (DOC) cycling". Frontiers in Marine Science, 7: 466. doi : 10.3389/fmars.2020.00466. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  8. Monroy, P., Hernández-García, E., Rossi, V. and López, C. (2017) "Modeling the dynamical sinking of biogenic particles in oceanic flow". Nonlinear Processes in Geophysics, 24(2): 293–305. doi:10.5194/npg-24-293-2017. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 3.0 International License.
  9. Simon, M., Grossart, H., Schweitzer, B. and Ploug, H. (2002) "Microbial ecology of organic aggregates in aquatic ecosystems". Aquatic microbial ecology, 28: 175–211. doi:10.3354/ame028175.
  10. Kirchman, David L.; Suzuki, Yoshimi; Garside, Christopher; Ducklow, Hugh W. (15 August 1991). "High turnover rates of dissolved organic carbon during a spring phytoplankton bloom". Nature. 352 (6336): 612–614. Bibcode:1991Natur.352..612K. doi:10.1038/352612a0. S2CID   4285758.
  11. Jaeckle, W.B.; Manahan, D.T. (1989). "Feeding by a "nonfeeding" larva: uptake of dissolved amino acids from seawater by lecithotrophic larvae of the gastropod Haliotis rufescens". Marine Biology. 103: 87–94. doi:10.1007/BF00391067. S2CID   84541307.
  12. Cheremisinoff, Nicholas; Davletshin, Anton (2015). "Hydraulic Fracturing Operations: Handbook of Environmental Management Practices". Environmental Management. ISBN   9781119099994.
  13. Elser, Stephen (2014). "Brown Water: The Ecological and Economic Implications of Increased Dissolved Organic Carbon in Lakes". Archived from the original on 25 September 2017.
  14. Wu, Qing; Zhao, Xin-Hua; Wang, Xiao-Dan (2008). "Relationship Between Heterotrophic Bacteria and Some Physical and Chemical Parameters in a Northern City's Drinking Water Distribution Networks of China". 2008 2nd International Conference on Bioinformatics and Biomedical Engineering. pp. 4713–4716. doi:10.1109/ICBBE.2008.336. ISBN   978-1-4244-1747-6. S2CID   24876521.
  15. "Dissolved Organic Carbon (DOC)".
  16. Narayana, P.S.; Varalakshmi, D; Pullaiah, T; Sambasiva Rao, K.R.S. (2018). Research Methodology in Zoology. Scientific Publishers. p. 225. ISBN   9789388172400.
  17. "Whatman glass microfiber filters, Grade GF/F". Merck.
  18. Knap, A. Michaels; A. Close; A. Ducklow; H. Dickson, A. (1994). Protocols for the Joint Global Ocean Flux studies (JGOFS) core measurements. JGOFS.
  19. Cauwet G (2002) "DOM in the Coastal Zone". In: Hansell D and Carlson C (Eds.) Biogeochemistry of Marine Dissolved Organic Matter, pages 579–610, Elsevier. ISBN   9780080500119.
  20. Tremblay, L. and Benner, R. (2006) "Microbial contributions to N-immobilization and organic matter preservation in decaying plant detritus". Geochimica et Cosmochimica Acta, 70(1): 133–146. doi : 10.1016/j.gca.2005.08.024.
  21. 1 2 Jiao, Nianzhi; Herndl, Gerhard J.; Hansell, Dennis A.; Benner, Ronald; Kattner, Gerhard; Wilhelm, Steven W.; Kirchman, David L.; Weinbauer, Markus G.; Luo, Tingwei; Chen, Feng; Azam, Farooq (2010). "Microbial production of recalcitrant dissolved organic matter: long-term carbon storage in the global ocean". Nature Reviews Microbiology. 8 (8): 593–599. doi:10.1038/nrmicro2386. PMID   20601964. S2CID   14616875.
  22. Lee, S.A., Kim, T.H. and Kim, G. (2020) "Tracing terrestrial versus marine sources of dissolved organic carbon in a coastal bay using stable carbon isotopes". Biogeosciences, 17(1). doi : 10.5194/bg-17-135-2020. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  23. 1 2 Vahatalo, A. V., Aarnos, H., and Mantyniemi, S. (2010). Biodegradability continuum and biodegradation kinetics of natural organic matter described by the beta distribution. Biogeochemistry 100, 227–240. doi: 10.1007/s10533-010-9419-4
  24. 1 2 3 4 Hansell, Dennis A. (2013). "Recalcitrant Dissolved Organic Carbon Fractions". Annual Review of Marine Science. 5: 421–445. doi:10.1146/annurev-marine-120710-100757. PMID   22881353.
  25. 1 2 Amon, R. M. W., and Benner, R. (1996). Bacterial utilization of different size classes of dissolved organic matter. Limnol. Oceanogr. 41, 41–51. doi: 10.4319/lo.1996.41.1.0041
  26. Benner, R., and Amon, R. M. (2015). The size-reactivity continuum of major bioelements in the ocean. Ann. Rev. Mar. Sci. 7, 185–205. doi: 10.1146/annurev-marine-010213-135126
  27. Thingstad, T. F., Havskum, H., Kaas, H., Nielsen, T. G., Riemann, B., Lefevre, D., et al. (1999). Bacteria-protist interactions and organic matter degradation under P-limited conditions: analysis of an enclosure experiment using a simple model. Limnol. Oceanogr. 44, 62–79. doi: 10.4319/lo.1999.44.1.0062
  28. Del-Giorgio, P., and Davies, J. (2003). "Patterns of dissolved organic matter lability and consumption across aquatic ecosystems", in Aquatic Ecosystems: Interactivity of Dissolved Organic Matter, eds S. E. G. Findlay and R. L. Sinsabaugh (San Diego, CA: Academic Press), 399–424. doi: 10.1016/B978-012256371-3/50018-4
  29. Bianchi, T. S. (2011). The role of terrestrially derived organic carbon in the coastal ocean: a changing paradigm and the priming effect. Proc. Natl. Acad. Sci. U.S.A. 108, 19473–19481. doi: 10.1073/pnas.1017982108
  30. Kattner, G., Simon, M., and Koch, B. P. (2011). "Molecular characterization of dissolved organic matter and constraints for prokaryotic utilization", in Microbial Carbon Pump in the Ocean, eds N. Jiao, F. Azam, and S. Sansers (Washington, DC: Science/AAAS).
  31. Keil, R. G., and Mayer, L. M. (2014). "Mineral matrices and organic matter", in Treatise on Geochemistry, 2nd Edn, eds H. Holland and K. Turekian (Oxford: Elsevier), 337–359. doi: 10.1016/B978-0-08-095975-7.01024-X
  32. Bianchi, T. S., Cui, X., Blair, N. E., Burdige, D. J., Eglinton, T. I., and Galy, V. (2018). Centers of organic carbon burial and oxidation at the land-ocean interface. Org. Geochem. 115, 138–155. doi: 10.1016/j.orggeochem.2017.09.008
  33. Ward, N. D., Keil, R. G., Medeiros, P. M., Brito, D. C., Cunha, A. C., Dittmar, T., et al. (2013). Degradation of terrestrially derived macromolecules in the Amazon River. Nat. Geosci. 6, 530–533. doi: 10.1038/ngeo1817
  34. Myers-Pigg, A. N., Louchouarn, P., Amon, R. M. W., Prokushkin, A., Pierce, K., and Rubtsov, A. (2015). Labile pyrogenic dissolved organic carbon in major Siberian Arctic rivers: implications for wildfire-stream metabolic linkages. Geophys. Res. Lett. 42, 377–385. doi: 10.1002/2014GL062762
  35. 1 2 3 Gmach, M.R., Cherubin, M.R., Kaiser, K. and Cerri, C.E.P. (2020) "Processes that influence dissolved organic matter in the soil: a review". Scientia Agricola, 77(3). doi:10.1590/1678-992x-2018-0164. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  36. 1 2 Shen, Y., Chapelle, F.H., Strom, E.W. and Benner, R. (2015) "Origins and bioavailability of dissolved organic matter in groundwater". Biogeochemistry, 122(1): 61–78. doi : 10.1038/s41467-019-11394-4. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  37. 1 2 Reitsema, R.E., Meire, P. and Schoelynck, J. (2018) "The future of freshwater macrophytes in a changing world: dissolved organic carbon quantity and quality and its interactions with macrophytes". Frontiers in plant science, 9: 629. doi:10.3389/fpls.2018.00629. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  38. 1 2 Thomas, J. D. (1997). The role of dissolved organic matter, particularly free amino acids and humic substances, in freshwater ecosystems. Freshw. Biol. 38, 1–36. doi: 10.1046/j.1365-2427.1997.00206.x
  39. Cole, J. J., Prairie, Y. T., Caraco, N. F., McDowell, W. H., Tranvik, L. J., Striegl, R. G., et al. (2007). Plumbing the global carbon cycle: integrating inland waters into the terrestrial carbon budget. Ecosystems 10, 172–185. doi: 10.1007/s10021-006-9013-8
  40. Raymond, P. A., Hartmann, J., Lauerwald, R., Sobek, S., McDonald, C., Hoover, M., et al. (2013). Global carbon dioxide emissions from inland waters. Nature 503, 355–359. doi: 10.1038/nature12760
  41. Kalbitz, K.; Solinger, S.; Park, J.H.; Michalzik, B.; Matzner, E. 2000. Controls on the dynamics of dissolved organic matter in soils: a review. Soil Science 165: 277–304.
  42. Zech, W.; Senesi, N.; Guggenberger, G.; Kaiser, K.; Lehmann, J.; Miano, T.M.; Miltner, A.; Schroth, G. 1997. Factors controlling humification and mineralization of soil organic matter in the tropics. Geoderma 79: 117–161.
  43. 1 2 Saidy, A.R.; Smernik, R.J.; Baldock, J.A.; Kaiser, K.; Sanderman, J. 2015. Microbial degradation of organic carbon sorbed to phyllosilicate clays with and without hydrous iron oxide coating. European Journal of Soil Science 66: 83–94.
  44. Kaiser, K.; Guggenberger, G. 2007. Sorptive stabilization of organic matter by microporous goethite: sorption into small pores vs. surface complexation. European Journal of Soil Science 58: 45–59.
  45. 1 2 Veum, K.S.; Goyne, K.W.; Motavalli, P.P.; Udawatta, R.P. 2009. Runoff and dissolved organic carbon loss from a paired-watershed study of three adjacent agricultural Watersheds. Agriculture, Ecosystems & Environment 130: 115–122.
  46. Sparling, G.; Chibnall, E.; Pronger, J.; Rutledge, S.; Wall, A.; Campbell, D.; Schipper, L. 2016. Estimates of annual leaching losses of dissolved organic carbon from pastures on Allophanic soils grazed by dairy cattle, Waikato, New Zealand. New Zealand Journal of Agricultural Research 59: 32–49.
  47. Sobek, S., Tranvik, L. J., Prairie, Y. T., Kortelainen, P., and Cole, J. J. (2007). Patterns and regulation of dissolved organic carbon: an analysis of 7,500 widely distributed lakes. Limnol. Oceanogr. 52, 1208–1219. doi: 10.4319/lo.2007.52.3.1208
  48. Stumm, W., and Morgan, J. J. (1996). Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. Environmental Science and Technology. New York: John Wiley & Sons, Inc.
  49. Madsen, T. V., and Sand-Jensen, K. (1991). Photosynthetic carbon assimilation in aquatic macrophytes. Aquat. Bot. 41, 5–40. doi: 10.1016/0304-3770(91)90037-6
  50. Regnier, P., Friedlingstein, P., Ciais, P., Mackenzie, F. T., Gruber, N., Janssens, I. A., et al. (2013). Anthropogenic perturbation of the carbon fluxes from land to ocean. Nat. Geosci. 6, 597–607. doi: 10.1038/ngeo1830
  51. Luyssaert, S., Abril, G., Andres, R., Bastviken, D., Bellassen, V., Bergamaschi, P., et al. (2012). The European land and inland water CO2, CO, CH4 and N2O balance between 2001 and 2005. Biogeosciences 9, 3357–3380. doi: 10.5194/bg-9-3357-2012
  52. 1 2 Kawasaki, N., and Benner, R. (2006). Bacterial release of dissolved organic matter during cell growth and decline: molecular origin and composition. Limnol. Oceanogr. 51, 2170–2180. doi: 10.4319/lo.2006.51.5.2170
  53. Lønborg, C., Álvarez-Salgado, X. A., Davidson, K., and Miller, A. E. J. (2009). Production of bioavailable and refractory dissolved organic matter by coastal heterotrophic microbial populations. Estuar. Coast. Shelf Sci. 82, 682–688. doi: 10.1016/j.ecss.2009.02.026
  54. Wada, S., Aoki, M. N., Tsuchiya, Y., Sato, T., Shinagawa, H., and Hama, T. (2007). Quantitative and qualitative analyses of dissolved organic matter released from Ecklonia cava Kjellman, in Oura Bay, Shimoda, Izu Peninsula, Japan. J. Exp. Mar. Biol. Ecol. 349, 344–358. doi: 10.1016/j.jembe.2007.05.024
  55. Willey, J. D., Kieber, R. J., Eyman, M. S. Jr., and Brooks Avery, G. (2000). Rainwater dissolved organic carbon concentrations and global flux. Glob. Biogeochem. Cycles 14, 139–148. doi: 10.1029/1999GB900036
  56. 1 2 Raymond, P. A., and Spencer, R. G. M. (2015). "Riverine DOM", in Biogeochemistry of Marine Dissolved Organic Matter, eds D. A. Hansell and C. A. Carlson (Amsterdam: Elsevier), 509–533. doi: 10.1016/B978-0-12-405940-5.00011-X
  57. Dachs, J., and Méjanelle, L. (2010). Organic pollutants in coastal waters, sediments, and biota: a relevant driver for ecosystems during the anthropocene? Estuarines Coasts 33, 1–14. doi: 10.1007/s12237-009-9255-8
  58. Hansell, Dennis; Carlson, Craig; Repeta, Daniel; Schlitzer, Reiner (2009). "Dissolved Organic Matter in the Ocean: A Controversy Stimulates New Insights". Oceanography. 22 (4): 202–211. doi:10.5670/oceanog.2009.109. hdl: 1912/3183 . S2CID   129511530.
  59. 1 2 3 4 Carlson, Craig A.; Hansell, Dennis A. (2015). "DOM Sources, Sinks, Reactivity, and Budgets". Biogeochemistry of Marine Dissolved Organic Matter. pp. 65–126. doi:10.1016/B978-0-12-405940-5.00003-0. ISBN   9780124059405.
  60. Zigah, Prosper K.; McNichol, Ann P.; Xu, Li; Johnson, Carl; Santinelli, Chiara; Karl, David M.; Repeta, Daniel J. (2017). "Allochthonous sources and dynamic cycling of ocean dissolved organic carbon revealed by carbon isotopes". Geophysical Research Letters. 44 (5): 2407–2415. Bibcode:2017GeoRL..44.2407Z. doi:10.1002/2016GL071348. hdl: 1912/8912 . S2CID   55057882.
  61. 1 2 3 Mentges, A.; Feenders, C.; Deutsch, C.; Blasius, B.; Dittmar, T. (2019). "Long-term stability of marine dissolved organic carbon emerges from a neutral network of compounds and microbes". Scientific Reports. 9 (1): 17780. Bibcode:2019NatSR...917780M. doi:10.1038/s41598-019-54290-z. PMC   6883037 . PMID   31780725. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  62. Karl, D. M., Hebel, D. V., Bjorkman, K., and Letelier, R. M. (1998). The role of dissolved organic matter release in the productivity of the oligotrophic north Pacific Ocean. Limnol. Oceanogr. 43, 1270–1286. doi: 10.4319/lo.1998.43.6.1270
  63. Wetz, M. S., and Wheeler, P. A. (2007). Release of dissolved organic matter by coastal diatoms. Limnol. Oceanogr. 52, 798–807. doi: 10.4319/lo.2007.52.2.0798
  64. Thornton, D. C. O. (2014). Dissolved organic matter (DOM) release by phytoplankton in the contemporary and future ocean. Eur. J. Phycol. 49, 20–46. doi: 10.1080/09670262.2013.875596
  65. Boekell, W. H. M. V., Hansen, F. C., Riegman, R., and Bak, R. P. M. (1992). Lysis-induced decline of a Phaeocystis spring bloom and coupling with the microbial foodweb. Mar. Ecol. Prog. Ser. 81, 269–276. doi: 10.3354/meps081269
  66. 1 2 Hygum, B. H., Petersen, J. W., and Søndergaard, M. (1997). Dissolved organic carbon released by zooplankton grazing activity- a high quality substrate pool for bacteria. J. Plankton Res. 19, 97–111. doi: 10.1093/plankt/19.1.97
  67. Lampert, W. (1978). Release of dissolved organic carbon by grazing zooplankton. Limnol. Oceanogr. 23, 831–834. doi: 10.4319/lo.1978.23.4.0831
  68. Jumars, P. A., Penry, D. L., Baross, J. A., and Perry, M. J. (1989). Closing the microbial loop: dissolved carbon pathway to heterotrophic bacteria from incomplete ingestion, digestion and absorption in animals. Deep Sea Res. 36, 483–495. doi: 10.1016/0198-0149(89)90001-0
  69. Iturriaga, R., and Zsolnay, A. (1981). Transformation of some dissolved organic compounds by a natural heterotrophic population. Mar. Biol. 62, 125–129. doi: 10.1007/BF00388174
  70. 1 2 Ogawa, H.; Amagai, Y.; Koike, I.; Kaiser, K.; Benner, R. (2001). "Production of refractory dissolved organic matter by bacteria". Science. 292 (5518): 917–920. Bibcode:2001Sci...292..917O. doi:10.1126/science.1057627. PMID   11340202. S2CID   36359472.
  71. McCarthy, M., Pratum, T., Hedges, J., and Benner, R. (1997). Chemical composition of dissolved organic nitrogen in the ocean. Nature 390, 150–154. doi: 10.1038/36535
  72. Suttle, C. A. (2005). Viruses in the sea. Nature 437, 356–361. doi: 10.1038/nature04160
  73. Weinbauer, M. A. G. (2004). Ecology of prokaryotic viruses. FEMS Microbiol. Rev. 28, 127–181. doi: 10.1016/j.femsre.2003.08.001
  74. Lønborg, C., Middelboe, M., and Brussaard, C. P. D. (2013). Viral lysis of Micromonas pusilla: impacts on dissolved organic matter production and composition. Biogeochemistry 116, 231–240. doi: 10.1007/s10533-013-9853-1
  75. Hansell DA and Craig AC (2015) "Marine Dissolved Organic Matter and the Carbon Cycle". Oceanography, 14(4): 41–49. doi : 10.5670/oceanog.2001.05. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  76. Krabberød, AK; Bjorbækmo, MFM; Shalchian-Tabrizi, K.; Logares, R. (2017). "Exploring the oceanic microeukaryotic interactome with metaomics approaches". Aquatic Microbial Ecology. 79: 1–12. doi: 10.3354/ame01811 . hdl: 10261/153315 . CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
  77. Delong, Edward F.; Karl, David M. (2005). "Genomic perspectives in microbial oceanography". Nature. 437 (7057): 336–342. Bibcode:2005Natur.437..336D. doi:10.1038/nature04157. PMID   16163343. S2CID   4400950.
  78. 1 2 3 4 5 Wagner, S., Schubotz, F., Kaiser, K., Hallmann, C., Waska, H., Rossel, P.E., Hansman, R., Elvert, M., Middelburg, J.J., Engel, A. and Blattmann, T.M. (2020) "Soothsaying DOM: A current perspective on the future of oceanic dissolved organic carbon". Frontiers in Marine Science, 7:341. doi : 10.3389/fmars.2020.00341. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  79. Brilinsky, M. (1977). Release of dissolved organic matter by some marine macrophytes. Mar. Biol. 39, 213–220. doi: 10.1007/BF00390995
  80. Pregnall, A. M. (1983). Release of dissolved organic carbon from the estuarine intertidal macroalga Enteromorpha prolifera. Mar. Biol. 73, 37–42. doi: 10.1007/BF00396283
  81. Penhale, P. A., and Smith, W. O. (1977). Excretion of dissolved organic carbon by eelgrass (Zostera marina) and its epiphytes. Limnol. Oceanogr. 22, 400–407. doi: 10.4319/lo.1977.22.3.0400
  82. 1 2 Barrón, C., and Duarte, C. M. (2015). Dissolved organic carbon pools and export from the coastal ocean. Glob. Biogeochem. Cycles 29, 1725–1738. doi: 10.1002/2014GB005056
  83. Martin, P., Cherukuru, N., Tan, A.S., Sanwlani, N., Mujahid, A. and Müller, M.(2018) "Distribution and cycling of terrigenous dissolved organic carbon in peatland-draining rivers and coastal waters of Sarawak, Borneo", Biogeosciences, 15(2): 6847–6865. doi : 10.5194/bg-15-6847-2018. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  84. Hewson, I., O'neil, J. M., Fuhrman, J. A., and Dennison, W. C. (2001). Virus-like particle distribution and abundance in sediments and overlying waters along eutrophication gradients in two subtropical estuaries. Limnol. Oceanogr. 46, 1734–1746. doi: 10.4319/lo.2001.46.7.1734
  85. Burdige, D. J., and Gardner, K. G. (1998). Molecular weight distribution of dissolved organic carbon in marine sediment pore waters. Mar. Chem. 62, 45–64. doi: 10.1016/S0304-4203(98)00035-8
  86. Burdige, D. J., and Komada, T. (2014). "Sediment pore waters", in Biogeochemistry of Marine Dissolved Organic Matter, eds D. A. Hansen and C. A. Carlson (Cambridge, MA: Academic Press), 535–577. doi: 10.1016/B978-0-12-405940-5.00012-1
  87. Komada, T., and Reimers, C. E. (2001). Resuspension-induced partitioning of organic carbon between solid and solution phases from a river–ocean transition. Mar. Chem. 76, 155–174. doi: 10.1016/S0304-4203(01)00055-X
  88. Dittmar, T., and Koch, B. P. (2006). Thermogenic organic matter dissolved in the abyssal ocean. Mar. Chem. 102, 208–217. doi: 10.1016/j.marchem.2006.04.003
  89. Dittmar, T., and Paeng, J. (2009). A heat-induced molecular signature in marine dissolved organic matter. Nat. Geosci. 2, 175–179. doi: 10.1038/ngeo440
  90. Burnett, W. C., Aggarwal, P. K., Aureli, A., Bokuniewicz, H., Cable, J. E., Charette, M. A., et al. (2006). Quantifying submarine groundwater discharge in the coastal zone via multiple methods. Sci. Total Environ. 367, 498–543. doi: 10.1016/j.scitotenv.2006.05.009
  91. Longnecker, K., and Kujawinski, E. B. (2011). Composition of dissolved organic matter in groundwater. Geochim. Cosmochim. Acta 75, 2752–2761. doi: 10.1016/j.gca.2011.02.020
  92. Webb, J. R., Santos, I. R., Maher, D. T., Tait, D. R., Cyronak, T., Sadat-Noori, M., et al. (2019). Groundwater as a source of dissolved organic matter to coastal waters: insights from radon and CDOM observations in 12 shallow coastal systems. Limnol. Oceanogr. 64, 182–196. doi: 10.1002/lno.11028
  93. 1 2 Lang, S. Q., Butterfield, D. A., Lilley, M. D., Paul Johnson, H., and Hedges, J. I. (2006). Dissolved organic carbon in ridge-axis and ridge-flank hydrothermal systems. Geochim. Cosmochim. Acta 70, 3830–3842. doi: 10.1016/j.gca.2006.04.031
  94. Kerner, M., Hohenberg, H., Ertl, S., Reckermann, M., and Spitzy, A. (2003). Self-organization of dissolved organic matter tomicelle-like microparticles in river water. Nature 422, 150–154. doi: 10.1038/nature01469
  95. Chin, W. C., Orellana, M. V., and Verdugo, P. (1998). Spontaneous assembly of marine dissolved organic matter into polymer gels. Nature 391, 568–572. doi: 10.1038/35345
  96. 1 2 3 Moran, M. A., and Zepp, R. G. (1997). Role of photoreactions in the formation of biologically labile compounds from dissolved organic matter. Limnol. Oceanogr. 42, 1307–1316. doi: 10.4319/lo.1997.42.6.1307
  97. 1 2 Mopper, K., Kieber, D. J., and Stubbins, A. (2015). "Marine photochemistry of organic matter", in Biogeochemistry of Marine Dissolved Organic Matter, eds C. A. Carlson and D. A. Hansell (Amsterdam: Elsevier), 389–450. doi: 10.1016/B978-0-12-405940-5.00008-X
  98. Lønborg, C., and Álvarez-Salgado, X. A. (2012). Recycling versus export of bioavailable dissolved organic matter in the coastal ocean and efficiency of the continental shelf pump. Glob. Biogeochem. Cycles 26:GB3018. doi: 10.1029/2012GB004353
  99. Carlson, C. A., and Hansell, D. A. (2015). "DOM sources, sinks, reactivity, and budgets", in Biogeochemistry of Marine Dissolved Organic Matter, eds C. A. Carlson and D. A. Hansell (San Diego, CA: Academic Press), 65–126. doi: 10.1016/B978-0-12-405940-5.00003-0
  100. 1 2 Shen, Yuan; Benner, Ronald (2018). "Mixing it up in the ocean carbon cycle and the removal of refractory dissolved organic carbon". Scientific Reports. 8 (1): 2542. Bibcode:2018NatSR...8.2542S. doi:10.1038/s41598-018-20857-5. PMC   5803198 . PMID   29416076. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  101. 1 2 Sholkovitz, E. R. (1976). Flocculation of dissolved organic and inorganic matter during the mixing of river water and seawater. Geochim. Cosmochim. Acta 40, 831–845. doi: 10.1016/0016-7037(76)90035-1
  102. Tranvik, L. J., and Sieburth, J. M. (1989). Effects of flocculated humic matter on free and attached pelagic microorganisms. Limnol. Oceanogr. 34, 688–699. doi: 10.4319/lo.1989.34.4.0688
  103. Mulholland, P. J. (1981). Formation of Particulate Organic Carbon in Water from a Southeastern Swamp-Stream. Limnol. Oceanogr. 26, 790–795. doi: 10.4319/lo.1981.26.4.0790
  104. Powell, R. T., Landing, W. M., and Bauer, J. E. (1996). Colloidal trace metals, organic carbon and nitrogen in a southeastern U.S. estuary. Mar. Chem. 55, 165–176. doi: 10.1016/S0304-4203(96)00054-0
  105. Sholkovitz, E. R., Boyle, E. A., and Price, N. B. (1978). The removal of dissolved humic acids and iron during estuarine mixing. Earth Planet. Sci. Lett. 40, 130–136. doi: 10.1016/0012-821X(78)90082-1
  106. Volk, C., Bell, K., Ibrahim, E., Verges, D., Amy, G., and Lechevallier, M. (2000). Impact of enhanced and optimized coagulation on removal of organic matter and its biodegradable fraction in drinking water. Water Res. 34, 3247–3257. doi: 10.1016/S0043-1354(00)00033-6
  107. Williamson, C. E., Stemberger, R. S., Morris, D. P., Frost, T. A., and Paulsen, S. G. (1996). Ultraviolet radiation in North American lakes: attenuation estimates from DOC measurements and implications for plankton communities. Limnol. Oceanogr. 41, 1024–1034. doi: 10.4319/lo.1996.41.5.1024
  108. Williamson, C. E., Overholt, E. P., Pilla, R. M., Leach, T. H., Brentrup, J. A., Knoll, L. B., et al. (2015). Ecological consequences of longterm browning in lakes. Sci. Rep. 5:18666. doi: 10.1038/srep18666
  109. Jeffrey, W. H., Aas, P., Lyons, M. M., Coffin, R. B., Pledger, R. J., and Mitchell, D. L. (1996). Ambient solar radiation-induced photodamage in marine bacterioplankton. Photochem. Photobiol. 64, 419–427. doi: 10.1111/j.1751-1097.1996.tb03086.x
  110. Rhode, S. C., Pawlowski, M., and Tollrian, R. (2001). The impact of ultraviolet radiation on the vertical distribution of zooplankton of the genus Daphnia. Nature 412, 69–72. doi: 10.1038/35083567
  111. Miller, W. L., and Zepp, R. G. (1995). Photochemical production of dissolved inorganic carbon from terrestrial organic matter: significance of the oceanic organic carbon cycle. Geophys. Res. Lett. 22, 417–420. doi: 10.1029/94GL03344
  112. Moran, M. A., Sheldon, W. M., and Zepp, R. G. (2000). Carbon loss and optical property changes during long-term photochemical and biological degradation of estuarine dissolved organic matter. Limnol. Oceanogr. 45, 1254–1264. doi: 10.4319/lo.2000.45.6.1254
  113. Kieber, R. J., Hydro, L. H., and Seaton, P. J. (1997). Photooxidation of triglycerides and fatty acids in seawater: implication toward the formation of marine humic substances. Limnol. Oceanogr. 42, 1454–1462. doi: 10.4319/lo.1997.42.6.1454
  114. Berto, S., Laurentiis, E. D., Tota, T., Chiavazza, E., Daniele, P. G., Minella, M., et al. (2016). Properties of the humic-like material arising from the phototransformation of L-tyrosine. Sci. Total Environ. 546, 434–444. doi: 10.1016/j.scitotenv.2015.12.047
  115. Hudson, J. J., Dillon, P. J., and Somers, K. M. (2003). Long-term patterns in dissolved organic carbon in boreal lakes: the role of incident radiation, precipitation, air temperature, southern oscillation and acid deposition. Hydrol. Earth Syst. Sci. 7, 390–398. doi: 10.5194/hess-7-390-2003
  116. Benner, R., Benitez-Nelson, B., Kaiser, K., and Amon, R. M. W. (2004). Export of young terrigenous dissolved organic carbon from rivers to the Arctic Ocean. Geophys. Res. Lett. 31:L05305. doi: 10.1029/2003GL019251
  117. Obernosterer, I., and Herndl, G. J. (1995). Phytoplankton extracellular release and bacterial growth: dependence on the inorganic N:P ratio. Mar. Ecol. Prog. Ser. 116, 247–257. doi: 10.3354/meps116247
  118. Benner, R., and Ziegler, S. (1999). "Do photochemical transformations of dissolved organic matter produce biorefractory as well as bioreactive substrates?" in Proceedings of the 8th International Symposium on Microbial Ecology, eds C. R. Bell, M. Brylinsky, and P. Johnson-Green (Port Aransas, TX: University of Texas at Austin).
  119. Sulzberger, B., and Durisch-Kaiser, E. (2009). Chemical characterization of dissolved organic matter (DOM): a prerequisite for understanding UV-induced changes of DOM absorption properties and bioavailability. Aquat. Sci. 71, 104–126. doi: 10.1007/s00027-008-8082-5
  120. Miller, W. L., and Moran, M. A. (1997). Interaction of photochemical and microbial processes in the degradation of refractory dissolved organic matter from a coastal marine environment. Limnol. Oceanogr. 42, 1317–1324. doi: 10.4319/lo.1997.42.6.1317
  121. Stubbins, A., Uher, G., Law, C. S., Mopper, K., Robinson, C., and Upstill-Goddard, R. C. (2006). Open-ocean carbon monoxide photoproduction. Deep Sea Res. II Top. Stud. Oceanogr. 53, 1695–1705. doi: 10.1016/j.dsr2.2006.05.011
  122. Miller, W. L., Moran, M. A., Sheldon, W. M., Zepp, R. G., and Opsahl, S. (2002). Determination of apparent quantum yield spectra for the formation of biologically labile photoproducts. Limnol. Oceanogr. 47, 343–352. doi: 10.4319/lo.2002.47.2.0343
  123. Andrews, S. S., and Zafiriou, O. C. (2000). Photochemical oxygen consumption in marine waters: a Major soink for colored dissolved organic matter? Limnol. Oceanogr. 45, 267–277. doi: 10.4319/lo.2000.45.2.0267
  124. Wang, X.-C., Chen, R. F., and Gardner, G. B. (2004). Sources and transport of dissolved and particulate organic carbon in the Mississippi River estuary and adjacent coastal waters of the northern Gulf of Mexico. Mar. Chem. 89, 241–256. doi: 10.1016/j.marchem.2004.02.014
  125. Walker, Brett D.; Beaupré, Steven R.; Guilderson, Thomas P.; McCarthy, Matthew D.; Druffel, Ellen R. M. (2016). "Pacific carbon cycling constrained by organic matter size, age and composition relationships". Nature Geoscience. 9 (12): 888–891. Bibcode:2016NatGe...9..888W. doi:10.1038/ngeo2830.
  126. Khatiwala, S.; Primeau, F.; Holzer, M. (2012). "Ventilation of the deep ocean constrained with tracer observations and implications for radiocarbon estimates of ideal mean age". Earth and Planetary Science Letters. 325–326: 116–125. Bibcode:2012E&PSL.325..116K. doi:10.1016/j.epsl.2012.01.038. S2CID   7017553.
  127. Follett, Christopher L.; Repeta, Daniel J.; Rothman, Daniel H.; Xu, Li; Santinelli, Chiara (2014). "Hidden cycle of dissolved organic carbon in the deep ocean". Proceedings of the National Academy of Sciences. 111 (47): 16706–16711. Bibcode:2014PNAS..11116706F. doi: 10.1073/pnas.1407445111 . PMC   4250131 . PMID   25385632.
  128. Sharp, Jonathan H. (6 August 1996). "Marine dissolved organic carbon: Are the older values correct?". Marine Chemistry. 56 (3–4): 265–277. doi:10.1016/S0304-4203(96)00075-8.
  129. Sondergaard, Morten; Mathias Middelboe (9 March 1995). "A cross-system analysis of labile dissolved organic carbon" (PDF). Marine Ecology Progress Series. 118: 283–294. Bibcode:1995MEPS..118..283S. doi: 10.3354/meps118283 .
  130. Gruber, David F.; Jean-Paul Simjouw; Sybil P. Seitzinger; Gary L. Taghon (June 2006). "Dynamics and Characterization of Refractory Dissolved Organic Matter Produced by a Pure Bacterial Culture in an Experimental Predator-Prey System". Applied and Environmental Microbiology. 72 (6): 4184–4191. Bibcode:2006ApEnM..72.4184G. doi:10.1128/AEM.02882-05. PMC   1489638 . PMID   16751530.
  131. 1 2 Hansell, Dennis A.; Craig A. Carlson; Daniel J. Repeta; Reiner Schlitzer (2009). "Dissolved Organic Matter in the Ocean: A Controversy Stimulates New Insights". Oceanography. 22 (4): 202–211. doi: 10.5670/oceanog.2009.109 . hdl: 1912/3183 .
  132. Stubbins, A., Niggemann, J., and Dittmar, T. (2012). Photo-lability of deep ocean dissolved black carbon. Biogeosciences 9, 1661–1670. doi: 10.5194/bg-9-1661-2012
  133. Hodson, R. E., Maccubbin, A. E., and Pomeroy, L. R. (1981). Dissolved adenosine triphosphate utilization by free-living and attached bacterioplankton. Mar. Biol. 64, 43–51. doi: 10.1007/bf00394079
  134. Hollibaugh, J. T., and Azam, F. (1983). Microbial degradation of dissolved proteins in seawater. Limnol. Oceanogr. 28, 1104–1116. doi: 10.4319/lo.1983.28.6.1104
  135. Ferguson, R. L., and Sunda, W. G. (1984). Utilization of amino acids by planktonic marine bacteria: importance of clean technique and low substrate additions. Limnol. Oceanogr. 29, 258–274. doi: 10.4319/lo.1984.29.2.0258
  136. Kaiser, K., and Benner, R. (2008). Major bacterial contribution to the ocean reservoir of detrital organic carbon and nitrogen. Limnol. Oceanogr. 53, 99–112. doi: 10.4319/lo.2008.53.1.0099
  137. Dittmar, Thorsten (2015). "Reasons Behind the Long-Term Stability of Dissolved Organic Matter". Biogeochemistry of Marine Dissolved Organic Matter. pp. 369–388. doi:10.1016/B978-0-12-405940-5.00007-8. ISBN   9780124059405.
  138. Jannasch, Holger W. (1967). "Growth of Marine Bacteria at Limiting Concentrations of Organic Carbon in Seawater1". Limnology and Oceanography. 12 (2): 264–271. Bibcode:1967LimOc..12..264J. doi: 10.4319/lo.1967.12.2.0264 .
  139. Arrieta, J. M.; Mayol, E.; Hansman, R. L.; Herndl, G. J.; Dittmar, T.; Duarte, C. M. (2015). "Dilution limits dissolved organic carbon utilization in the deep ocean". Science. 348 (6232): 331–333. Bibcode:2015Sci...348..331A. doi: 10.1126/science.1258955 . PMID   25883355. S2CID   28514618.
  140. Traving, Sachia J.; Thygesen, Uffe H.; Riemann, Lasse; Stedmon, Colin A. (2015). "A Model of Extracellular Enzymes in Free-Living Microbes: Which Strategy Pays Off?". Applied and Environmental Microbiology. 81 (21): 7385–7393. Bibcode:2015ApEnM..81.7385T. doi:10.1128/AEM.02070-15. PMC   4592861 . PMID   26253668.
  141. 1 2 Nebbioso, Antonio; Piccolo, Alessandro (January 2013). "Molecular characterization of dissolved organic matter (DOM): a critical review". Analytical and Bioanalytical Chemistry. 405 (1): 109–124. doi:10.1007/s00216-012-6363-2. ISSN   1618-2642. PMID   22965531. S2CID   36714947.
  142. 1 2 Minor, Elizabeth C.; Swenson, Michael M.; Mattson, Bruce M.; Oyler, Alan R. (21 August 2014). "Structural characterization of dissolved organic matter: a review of current techniques for isolation and analysis". Environmental Science: Processes & Impacts. 16 (9): 2064–2079. doi:10.1039/C4EM00062E. PMID   24668418.
  143. Green, Nelson W.; Perdue, E. Michael; Aiken, George R.; Butler, Kenna D.; Chen, Hongmei; Dittmar, Thorsten; Niggemann, Jutta; Stubbins, Aron (20 April 2014). "An intercomparison of three methods for the large-scale isolation of oceanic dissolved organic matter". Marine Chemistry. 161: 14–19. Bibcode:2014MarCh.161...14G. doi:10.1016/j.marchem.2014.01.012. ISSN   0304-4203.