Skip to content
BY 4.0 license Open Access Published by De Gruyter December 30, 2019

Magmatic carbon outgassing and uptake of CO2 by alkaline waters

  • Marie Edmonds EMAIL logo , Benjamin Tutolo , Kayla Iacovino and Yves Moussallam
From the journal American Mineralogist

Abstract

Much of Earth’s carbon resides in the “deep” realms of our planet: sediments, crust, mantle, and core. The interaction of these deep reservoirs of carbon with the surface reservoir (atmosphere and oceans) leads to a habitable surface environment, with an equitable atmospheric composition and comfortable range in temperature that together have allowed life to proliferate. The Earth in Five Reactions project (part of the Deep Carbon Observatory program) identified the most important carbon-bearing reactions of our planet, defined as those which perhaps make our planet unique among those in our Solar System, to highlight and review how the deep and surface carbon cycles connect. Here we review the important reactions that control the concentration of carbon dioxide in our atmosphere: outgassing from magmas during volcanic eruptions and during magmatic activity; and uptake of CO2 by alkaline surface waters. We describe the state of our knowledge about these reactions and their controls, the extent to which we understand the mass budgets of carbon that are mediated by these reactions, and finally, the implications of these reactions for understanding present-day climate change that is driven by anthropogenic emission of CO2.

Introduction

We present this paper as part of the “Earth in Five Reactions” series, which identifies the most important carbon-bearing reactions in Earth from the perspective of the “deep” carbon cycle. The concentration of CO2 in our atmosphere controls the habitability of our planet via our equitable climate and the existence of photosynthesizing organisms. How does the deep carbon cycle impact the concentration of CO2 in our atmosphere?

The Earth’s “surface reservoir,” i.e., the atmosphere, oceans, and sediments combined, of carbon makes up only a small fraction of the Earth’s total carbon budget, with most of the carbon inventory of the planet (estimates vary from 75 to 99%) residing in the Earth’s core (Wood 1993; Dasgupta and Hirschmann 2010a; Dasgupta et al. 2013; DePaolo 2015; Li et al. 2019). Over long geological timescales, fluxes of carbon from the mantle to the atmosphere and oceans via magmatism, volcanism, and metamorphic decarbonation, modulate the composition of Earth’s atmosphere and climate (Sleep and Zahnle 2001; Dasgupta 2013) (Fig. 1). The delicate carbon balance of the surface reservoir is largely maintained (Berner and Caldeira 1997), despite the immense size of Earth’s internal reservoirs and fluxes, which requires equally important fluxes of carbon back into the mantle, through subduction (Dasgupta 2013; Kelemen and Manning 2015; Hirschmann 2018). This return flux replenishes the carbon lost from the mantle via melting and degassing (Dasgupta 2013; Hirschmann 2018).

Figure 1 Simplified box model to show the carbon reservoirs and fluxes of Earth (excluding the core). The size of each reservoir is approximately indicated by the horizontal dimension of the box; the approximate residence time for carbon in each reservoir by the vertical dimension (Sundquist 1993; Dasgupta and Hirschmann 2010b; Friedlingstein et al. 2010; Burton et al. 2013; DePaolo 2015; Lee et al. 2019; Werner et al. 2019). The types of fluxes between each reservoir are shown against the arrows; the size of these fluxes is not indicated.
Figure 1

Simplified box model to show the carbon reservoirs and fluxes of Earth (excluding the core). The size of each reservoir is approximately indicated by the horizontal dimension of the box; the approximate residence time for carbon in each reservoir by the vertical dimension (Sundquist 1993; Dasgupta and Hirschmann 2010b; Friedlingstein et al. 2010; Burton et al. 2013; DePaolo 2015; Lee et al. 2019; Werner et al. 2019). The types of fluxes between each reservoir are shown against the arrows; the size of these fluxes is not indicated.

The volcanic, tectonic, and metamorphic flux of CO2 from Earth’s interior is large. For the volcanic portion, current estimates range from 280 to 360 Tg CO2 per year (equivalent to 6.4 to 8.2 × 1012 mol CO2 per year) (Burton et al. 2013; Gerlach 2011; Werner et al. 2019). Volcanic C is derived from the outgassing of silicate melts at low pressure. The tectonic flux includes direct outgassing of mantle-derived CO2 fluids (Chiodini et al. 2010; Mörner and Etiope 2002), as well as metamorphic decarbonation in orogenic belts (Kerrick and Caldeira 1998). Degassing of crustal- and mantle-derived fluids associated with extensional tectonic regimes (e.g., outgassing from continental rifts) constitutes a large fraction of the total outgassing flux; a recent estimate (which is also included in the “volcanic” estimate above), is hundreds of Tg per year of CO2 (Tamburello et al. 2018).

The majority of the volcanic flux is derived from the degassing of carbon from silicate melts generated in the mantle, linked to mantle plumes or plate boundary processes, which ascend toward the surface of the Earth. It has been shown that the mantle is heterogeneous with respect to carbon (Saal et al. 2002; Hirschmann and Dasgupta 2009; Hauri et al. 2018, 2019), which leads to melts produced in different settings containing different amounts of carbon. Mid-ocean ridge melts are largely carbon-depleted relative to those formed in arcs and hotspot settings (Blundy et al. 2010; Hauri et al. 2019). Carbon solubility in silicate melts falls with decreasing pressure, which causes melts to outgas their carbon almost entirely once they reach the shallow crust. This reaction might be represented by:

(1)CO2silicatemeltCO2g

and

(2)CO3silicatemelt2CO2g+Osilicatemelt2

for dissolved molecular CO2 and dissolved carbonate, respectively (where O2– is a non-bridging oxygen) (Guillot and Sator 2011). Both species are present in silicate melts, with molecular CO2 becoming more important as the silica content of the melt increases (and temperature decreases) (Stolper and Holloway 1988). The controls on silicate melt degassing (reactions 1 and 2) are explored further below. Melts may also assimilate carbon from the mantle lithosphere, or from the crust (Lee et al. 2013; Carter and Dasgupta 2015; Mason et al. 2017), where carbon resides as carbonate and organic carbon, and transfer this carbon to the atmosphere. Because of the low solubility of carbon in silicate melts, intrusive magmatism contributes to the outgassing flux of carbon, leading to diffuse emissions of CO2 in volcanic regions, often mediated by large-scale faults and fractures (Allard et al. 1991; Chiodini et al. 1995; Farrar et al. 1995; Hunt et al. 2017).

Volcanoes and extensional tectonic regions pump CO2 into the atmosphere (Fig. 1), triggering a systematic response in the surface reservoir (Berner 1999). An equilibrium exists between the CO2 in the atmosphere and in the oceans. On short timescales, the atmosphere-ocean system responds to an increase in atmospheric CO2 by allowing more carbon to dissolve in the ocean:

(3)CO2gasCO2aq

Aqueous CO2 dissociates to carbonate and bicarbonate, with the relative proportions dependent on the pH of the ocean water:

(4)CO2+H2O=HCO3+H+=C32+2H+

Adding more CO2 to the ocean increases acidity (at the expense of carbonate); increasing the alkalinity of the ocean will tend to draw down CO2. The residence time of carbon in the atmosphere-ocean-sediment system is relatively short; only ~100 ka (Berner 1994). Carbon is removed from the surface reservoir through weathering and removal of bicarbonate through rivers into the ocean. Then, carbon is precipitated from the ocean water as carbonate above the carbonate compensation depth (CCD), forming limestones on the ocean floor and carbonate in oceanic crust, which may eventually be returned to the mantle via subduction. A fraction of this carbon may be removed from the slab during subduction, via dissolution or metamorphic decarbonation (Kerrick and Connolly 2001; Gorman et al. 2006; Ague and Nicolescu 2014; Facq et al. 2014), migrating back to the surface in the forearc or subarc in the form of CO2-rich fluids and by dissolution within primary arc melts and transport to the surface through volcanism (Dutkiewicz et al. 2018).

The transfer of CO2 between gas and water and silicate melts therefore plays an outsized role in connecting its various “spheres.” Metamorphic and magmatic CO2 inputs from Earth’s interior enter the atmosphere through the forward expression of reactions 1 and 2, dissolve into its hydrosphere through the forward expression of reaction 3, and ultimately interact with cations derived from lithospheric weathering to form carbonate minerals, often through biosphere-mediated processes (Garrels 1983). Indeed, this silicate weathering feedback is the primary reason that the surface of our planet is clement (Stewart et al. 2019). Moreover, much of the carbon available to biological metabolisms such as photosynthesis (see Sheik et al. in revision) ultimately originates through a coupling of reactions 1+2 and 3. Due to their significance in connecting Earth’s spheres, reactions 1 (and 2) and 3 together were chosen as one of the five most important reactions that make Earth unique in the solar system by delegates of the Deep Carbon Observatory’s Earth in Five Reactions workshop.

In this paper, we review the current state of understanding of both reactions in detail, with application to the case of alkaline melts and fluids generated in continental rifts, which have had considerable importance for supplying CO2 fluxes to the surface reservoir through supercontinent cycles of the past (Brune et al. 2017) and in the modern day (Lee et al. 2016; Hunt et al. 2017; Werner et al. 2019). We discuss the dependence of these reactions on other parameters in the geological system, and we consider their importance for the Earth’s deep carbon cycle, tracking carbon reactions from the mantle to the surface. Finally, we draw attention to the relevance of both of these naturally occurring reactions, which have played such an important role in shaping Earth’s climate and surface environment over its long history, to our present-day societal challenges related to anthropogenic CO2 release and ocean drawdown of CO2.

Carbon degassing from silicate melts

Carbon exists in many forms in the Earth’s mantle: as fluids, dissolved as oxidized carbonate species (Pan and Galli 2016) or as reduced hydrocarbons (Sverjensky and Huang 2015) or as solid carbonate (Boulard et al. 2015; Cerantola et al. 2019); as elemental carbon (graphite and diamond; Shirey et al. 2013) in the deep mantle; and as Fe-metal carbide in the mantle (McCammon et al. 2020, this issue) core (Li and Fei 2003; Wood et al. 2013). Carbonated silicate melts may be ubiquitous in the mantle and may be the precursor to mid-ocean ridge basalts, kimberlites, and melts derived from subducting slabs (Dasgupta and Hirschmann 2010a; Moussallam et al. 2015). As silicate melts generated in the mantle ascend into the crust, they transport their dissolved carbon load toward the surface of the Earth. As they decompress, carbon dioxide exsolves and forms a separated supercritical fluid, transforming to a gas phase close to the surface. In this way, carbon may be transferred from the interior of the Earth to the surface; this is a critical flux of the deep carbon cycle (Fig. 1).

Because of the relatively high pressures in the crust and upper mantle (at depths of perhaps up to 40–50 km) under which silicate melt degassing occurs, generated CO2-rich fluid may migrate semi-independently of the magma through the crust, leading to large areas of so-called “diffuse” degassing at the surface (Chiodini et al. 2004), with the fluid making use of fractures in the crust to reach the atmosphere. Diffuse degassing takes place in a range of tectonic settings, e.g., at mid-ocean ridges, volcanic arcs over subduction zones, and submarine arc vents, in continental rifts (Hunt et al. 2017), and in large silicic calderas (Werner and Brantley 2003). In this way, the carbon content of unerupted magmas may yet reach the atmosphere, despite the source magma stalling in the crust. In the case of volcanic eruptions, CO2 is carried directly to the surface in ascending magmas (in both dissolved and exsolved forms), and is outgassed to the atmosphere either prior to or during eruption. In this section, we review our state of understanding of carbon degassing from silicate melts.

Controls on CO2 solubility in silicate melts

CO2 exists as both dissolved carbonate and as molecular CO2 in silicate glasses (Mysen et al. 1975; Fine and Stolper 1985; Fine and Stolper 1986). In basalts, basanites, and other silica-undersaturated magmas, carbon exists solely as carbonate (Blank and Brooker 1994; Moussallam et al. 2016), whereas in rhyolitic glasses, only molecular CO2 is found (Fogel and Rutherford 1990). Intermediate magmas and evolved, silica-undersaturated glasses such as phonolites, have both species present (Behrens et al. 2004). Carbon speciation in silicate melts is a strong function not only of melt silica content but also of melt structure and availability of cations such as Na+, Ca2+, and Mg2+ (Fine and Stolper 1985). The speciation of carbon dioxide in glasses (i.e., at surface P-T conditions) however is not necessarily that of the melt (i.e., at depth). Evidence from annealing experiments (Morizet et al. 2001; Nowak et al. 2003), molecular dynamics simulation studies (Guillot and Sator 2011; Morizet et al. 2015; Moussallam et al. 2016) and in situ spectroscopy studies (Konschak and Keppler 2014) show that the CO2 + O2– = CO32– equilibrium in the melt shifts toward CO2 with increasing temperature and that in the melt, molecular CO2 is present even in basaltic and carbonated melt compositions.

CO2 solubility in silicate melts is controlled by the melt composition via the attractive-repulsive behavior of each component with CO2 (Papale et al. 2006). Of these, Si and Al are by far the dominant factors, with CO2 solubility decreasing drastically with increasing Si and Al and increasing with increasing Ca, K, and Na content (plus more minor effect with increasing Mg and Fe; Fig. 2a) (Blank and Brooker 1994; Dixon et al. 1997; Shishkina et al. 2014; Moussallam et al. 2015). This is illustrated by the empirical relationship between CO2 solubility and the compositional “Pi parameter” (π, Dixon et al. 1997), which incorporates negative Si and Al terms and positive Ca, K, Na, Mg, and Fe terms.

Figure 2 Summary diagram to illustrate the major controls (melt composition and pressure) on CO2 solubility in silicate melts. (a) Compositionally dependent π (Pi) parameter and melt CO2 content at saturation for 50 synthetic basalt to alkali basalt melt compositions (colored symbols). CO2 concentrations calculated at 4 saturation pressures using MagmaSat (Ghiorso and Gualda 2015) in the ENKI framework (http://enki-portal.org/). Colored curves are best-fit to the data at each pressure and illustrate the π parameter as a good predictor of CO2 solubility. (b) Natural basalt melt CO2 and H2O contents (gray dots, data from GEOROC) plotted with isobars (colored curves; calculated for an alkali basalt using MagmaSat in the ENKI framework), which represent the saturation surface at given pressures.
Figure 2

Summary diagram to illustrate the major controls (melt composition and pressure) on CO2 solubility in silicate melts. (a) Compositionally dependent π (Pi) parameter and melt CO2 content at saturation for 50 synthetic basalt to alkali basalt melt compositions (colored symbols). CO2 concentrations calculated at 4 saturation pressures using MagmaSat (Ghiorso and Gualda 2015) in the ENKI framework (http://enki-portal.org/). Colored curves are best-fit to the data at each pressure and illustrate the π parameter as a good predictor of CO2 solubility. (b) Natural basalt melt CO2 and H2O contents (gray dots, data from GEOROC) plotted with isobars (colored curves; calculated for an alkali basalt using MagmaSat in the ENKI framework), which represent the saturation surface at given pressures.

The relatively dry and shallow melting regime beneath rift margins relative to other plate boundaries results in the generation very low-degree (<3%) partial mantle melts. Such melts are characteristically low in Si and rich in Ca, K, Na, and Mg resulting in a stronger propensity to dissolve available carbon (as carbonate). At mantle pressures, primary melts may dissolve several weight percent of carbon (up to 6 wt% at mantle pressures; Fig. 2a).

The solubility of CO2 increases with pressure for all silicate melt compositions (Fig. 2b). At 300 MPa in the crust, for example, basalts may hold 3000–4000 ppm in solution, depending on the amount of other volatile species present, decreasing to only a couple of 100 ppm at 50 MPa. Much of the CO2 load of a basaltic melt, which may reach 0.5 or even 1 wt% in the crust, therefore resides in the exsolved magmatic volatile phase in the mid- and upper-crust.

The presence of other volatiles affects the solubility of CO2 in silicate melts. Henry’s Law applies to volatile solubility at pressures pertaining to the mid- and upper-crust. For a vapor-saturated melt, where more than one volatile species is present, the concentration, or activity, of a particular species is diluted, thereby reducing its solubility in the melt (Moore 2008). The magnitude of this effect on melt CO2 solubility is illustrated in Figure 2b, whereas the co-existing vapor phase becomes more H2O-rich, the concentration (solubility) of CO2 in the melt is reduced (Dixon and Stolper 1995; Dixon et al. 1995; Moore et al. 1998; Newman and Lowenstern 2002; Moore 2008). This occurs in nature during a process known as second boiling, where magmas crystallize at a fixed pressure in the crust (Candela 1997). During crystallization, the melt volume reduces, and volatiles are forced out of solution and outgassed. Over time, the vapor phase thus produced gets progressively more water-rich, thereby reducing the solubility of CO2 in the remaining melt, as observed in quartz melt inclusions erupted in the Bishop Tuff, for example (Wallace et al. 1999). Conversely, underplating mafic magmas producing exsolved CO2 may “flush” overlying stored magmas (Cashman 2004; Blundy et al. 2010; Caricchi et al. 2018). In this case, the addition of CO2 raises the concentration of CO2 in the coexisting vapor phase, thereby enhancing CO2 solubility in the silicate melt. This process of CO2 flushing has been invoked to explain anomalously high CO2 concentrations in olivine-hosted melt inclusions from Etna, for example (Métrich and Wallace 2008).

The volcanic and tectonic flux of CO2 to the surface environment encompasses contributions from mid-ocean ridges (which is released dominantly into the oceans), from continental rifts, volcanic arcs, hotspots, extensional regions where the Earth’s mantle supplies CO2-rich fluids direct to the surface, and orogenic belts where limestones are heated and devolatilized (Sleep and Zahnle 2001). The fraction of this CO2 that is released directly into the atmosphere, and that which is absorbed into water bodies such as oceans, lakes, and groundwater aquifers, is unknown. It has been well documented that a large fraction of the CO2 released in extensional regions are trapped by aquifer systems, which, in turn, allow quantification of the CO2 produced over large areas, e.g., in the eastern Himalayas (Evans et al. 2008) and in the Apennines region of Italy (Chiodini et al. 2000).

Uptake of magmatic CO2 by surface water bodies

Alkalinity and carbon storage in surface water bodies

Earth’s surface water bodies can play a vital role in controlling the ultimate fate of carbon degassed from silicate melts. The amount of inorganic carbon in a solution in equilibrium with gaseous CO2 is a function of the solution’s alkalinity and its pH (Fig. 3). In essence, alkalinity is a charge-balance equation defined in relation to the speciation of dissolved inorganic carbon. Formally, carbonate alkalinity is defined (Stumm and Morgan 1996) as:

Figure 3 Diagram showing the relationship between pH, alkalinity, and total carbonate carbon concentration (after Deffeyes 1965); calculated by speciating solutions along axes using the Geochemist’s Workbench (Bethke et al. 2018). Labels on lines indicate the pH of the solution, and circles are analyses of East African Rift lakes collected by (Cerling 1979). Inset is plotted to emphasize the magnitude of the difference in carbon and alkalinity concentrations between seawater and alkaline rift lakes.
Figure 3

Diagram showing the relationship between pH, alkalinity, and total carbonate carbon concentration (after Deffeyes 1965); calculated by speciating solutions along axes using the Geochemist’s Workbench (Bethke et al. 2018). Labels on lines indicate the pH of the solution, and circles are analyses of East African Rift lakes collected by (Cerling 1979). Inset is plotted to emphasize the magnitude of the difference in carbon and alkalinity concentrations between seawater and alkaline rift lakes.

(5)Alkalinity=HCO3+2CO32+OHH+

where the brackets indicate concentrations of the bracketed species in mole-based units (e.g., molal).

Inorganic carbon enters solution through reaction 3, and, depending on the partial pressure of CO2 and the alkalinity and pH of the solution, is speciated into carbonic acid (CO2(aq)), bicarbonate (HCO3-), and carbonate (CO32–). Examination of Equation 5 shows that dissolving/exsolving CO2(g), an electrically neutral molecule, into/out of solution cannot directly impact alkalinity, although it may have the side effect of making minerals more or less soluble, which will, in turn, impact the solution alkalinity. Equation 5 shows that the total inorganic carbon in solution can be directly related to pH and alkalinity (Fig. 3). Nonetheless, meaningful use of Equation 4 in the study of dynamic aqueous environments begs the question—what are the concentrations of cations other than HCO3-, CO32–, H+, and OH-, and how might they contribute to the overall behavior (i.e., pH, carbon storage capacity) of the solution and/or reflect its history? Basalt has an excess of cations over CO2. Atmospheric CO2 binds with the cations once the rock weathers, making the weathering of basalt a carbon sink. For this purpose, an “alternative definition” of alkalinity reflecting the non-carbonate carbon portion of the charge balance (Stumm and Morgan 1996) is:

(6)Alkalinity=Na++K++2Ca2++2Mg2+Cl2SO422NO32.

In Equation 6, the impact of water chemistry on alkalinity and hence carbon storage and pH (Fig. 3) can be more readily discerned. Notably, this is a functional version of the equation that includes the “major” cationic and anionic species representative of most natural waters; it would need to be expanded for waters where other ions may be important. Equation 6 shows that the alkalinity is reduced by concentrations of anions such as Cl- and SO4, two of the dominant anions in seawater.

Aqueous carbon storage at rifted margins

At the same time as our understanding of carbon fluxes from continental rifts has increased through the efforts of DCO members and others (Lee et al. 2016; Brune et al. 2017; Hunt et al. 2017), attention has also been refocused on the fluxes of CO2 from mid-ocean ridges (e.g., Lund et al. 2016) and their variability in relation to long-term geologic variations. In the case of mid-ocean ridge systems, much of the CO2 degassing from the system is rapidly absorbed by the overlying seawater due to its high solubility at seafloor depths. For example, although CO2 concentrations of continuously venting hydrothermal fluids are commonly in the tens to hundreds of millimolal (e.g., Lilley et al. 2003), it is rare, although not unheard of to observe free-phase CO2 in the vicinity of submarine hydrothermal systems (Lupton et al. 2008).

It is now clear that continental rifts are important sites of magmatic CO2 release (Hunt et al. 2017): the eastern branch of the East African Rift, for example, may be emitting more CO2 (~3.9–32.7 Mt/year) than all of the world’s mid-ocean ridges combined (Lee et al. 2016; Hunt et al. 2017; Hauri et al. 2019), although there are large uncertainties attached to these estimates. Sites of continental breakup tap directly the vast volatile storage region in the sub-continental lithosphere, where small fraction, volatile-rich melts freeze as they cross their solidus as they ascend from the convecting mantle (Foley and Fischer 2017). Over geological time, supercontinent breakup may have released enormous fluxes of CO2 held both in the sub-continental lithosphere but also in the form of sediments on the continents, which may have played an important role in shaping Earth’s climate (Lee et al. 2013; Brune et al. 2017). Today, the CO2 emitted from the East African Rift may vary along-rift (Hunt et al. 2017). The CO2 migrates to the surface along faults at the rift margins (Hutchison et al. 2015), where it may be intercepted by surface water in the form of rift lakes.

The geology of the watersheds feeding rift lakes exerts an important control on the aqueous geochemistry of the lake waters (Cerling 1979, 1994; Wright 2012). Analyses presented by Cerling (1994) demonstrates that weathering of the volcanic rocks feeding East African Rift-hosted lakes provides mainly Na+, Ca2+, HCO3-, and relatively little Cl- and SO4 to the lake waters. When examined in the alkalinity framework represented by Equation 5, one can see the alkalinity of these lakes tends to be quite high (Fig. 3). If the lakes are in a closed basin, evaporation-dominated setting, then the alkalinity, and hence pH, tends to increase during evaporation (Garrels and MacKenzie 1967; Tutolo and Tosca 2018). As both pH and alkalinity increase, the amount of inorganic carbon stored in the lake also increases (Fig. 3), such that evaporation-dominated rift basin lake waters can contain concentrations of inorganic carbon significantly greater than seawater. This inorganic carbon can be deposited as carbonate minerals or converted to biomass and stored in lake sediments, often through interactions with cyanobacterial or algal communities (Wright 2012). Hence, a significant amount of carbon emitted from silicate melts in continental rifts may not go much further than the sediments deposited in the overlying lakes. Because the rift environment is dynamic, with lakes changing in size and alkalinity over time, intimately linked to the tectonics of rifting, the system does not necessarily reach a steady-state equilibrium; and the total carbon sequestered by lakes in rift environments remains unquantified. In this way, rifted margins are locations where the lithosphere, in the form of magmatic degassing, the hydrosphere, in the form of rift-basin lakes or seawater, and the biosphere, in the form of cyanobacteria, algae, and higher-order organisms, all interact with one another.

Implications for present-day anthropogenic CO2 release and ocean response

We have shown that the flux of carbon from Earth’s interior, and the potential for uptake of CO2 by alkaline surface waters, are first-order controls on the concentration of CO2 in our atmosphere. Up to now, we have confined our discussion to the natural cycles of magmatism and tectonics that have characterized our planet for the past two billion years. However, the processes that we discuss here are also relevant to a pressing predicament facing humankind in the 21st century. Namely, is anthropogenic CO2 generation a stronger driver of climate than CO2 injected into the atmosphere by natural processes? And how is the atmosphere-ocean system responding?

The answer to the first is unequivocally yes. Numerous estimates of CO2 delivered into the atmosphere by volcanism lead to an inescapable conclusion: anthropogenic sources of CO2 far exceed the volcanic and tectonic flux of CO2 into the atmosphere.

The burning of fossil fuels and the manufacture of cement released 31 000 ± 1845 Tg CO2 into the atmosphere in 2009 (Friedlingstein et al. 2010). Volcanic regions annually supply 280–360 Tg CO2 per year to the atmosphere and ocean system (Gerlach 2011; Burton et al. 2013; Werner et al. 2019), that is 83 to 137 times less than anthropogenic sources. The massive fluxes of carbon into the surface reservoir caused by anthropogenic activities, principally the burning of fossil fuels, has exceeded the Earth system’s ability to compensate, upsetting the delicate balance in the geologic carbon cycle (Berner and Caldeira 1997).

The ocean has absorbed around a third of the CO2 produced anthropogenically since the mid-20th century (Sabine et al. 2004) through Equation 3. Uptake of CO2 in ocean water decreases the pH by:

(7)CO2aq+H2OH2CO3HCO3+H+CO32+2H+

thereby increasing CO2 and bicarbonate at the expense of carbonate. Since the mid-20th century, it is estimated that ocean pH has dropped by 0.1 log units, which represents an almost 30% increase in H+ ions (Caldeira and Wickett 2003). The increase in acidity lowers the saturation state of calcite and aragonite in the oceans and is slowing or shutting down calcification in many marine organisms, including those that inhabit coral reefs (Orr et al. 2005; Zeebe et al. 2008; Doney et al. 2009).

Recent work has attempted to compare the recent rates of anthropogenic CO2 release to periods of dramatic carbon release from deep reservoirs to the atmosphere-ocean system in Earth’s history, which are recorded as light carbon isotope excursions in sedimentary records (Zeebe et al. 2016). A large carbon isotopic excursion and associated mass extinction at the end of the Permian has been linked to the outpouring of 7–15 million km3 basalt (Saunders 2005; Reichow et al. 2009; Black et al. 2012; Black and Gibson 2019) as well as the metamorphic devolatilization (Svensen et al. 2009) and combustion (Ogden and Sleep 2012) of buried coal by sills. This event may have released 20 000 to 30 000 Pg C over as long as 105 years, but probably in discrete pulses over much shorter timescales (Black et al. 2018). The fastest large carbon release of the Cenozoic (past 66 Myr) occurred at the onset of the Palaeocene–Eocene Thermal Maximum (~56 Myr ago) (Zachos et al. 2005), when 2500 to 4500 Pg carbon was emitted to the ocean-atmosphere system over up to 4000 yr (Bowen et al. 2015; Zeebe et al. 2016). These estimates yield annual carbon emissions during the PETM perhaps as little as 1 Pg carbon per year (Stocker et al. 2013; Zeebe et al. 2016) or as much as ~9.5 Pg carbon per year, which is similar to present-day anthropogenic fluxes (Bowen et al. 2015). The PETM was associated with massive extinction of benthic foraminifera, synchronous with a large decline in oceanic productivity and oxygenation (Winguth et al. 2012). CO2 released from combustion of fossil fuels in our atmosphere equilibrates, as discussed above, among the carbon reservoirs of the ocean, atmosphere, and biosphere over timescales of centuries. Geological sequestration of the CO2 from the surface reservoir; however, takes much longer: weathering processes and deposition of carbonate in oceanic crust occurs on timescales of 105–106 years (Lenton and Britton 2006; Archer et al. 2009).


Orcid 0000-0003-1243-137X

Open access: Article available to all readers online. This article is CC-BY.

† Special collection papers can be found online at http://www.minsocam.org/MSA/AmMin/special-collections.html


Acknowledgments and Funding

We acknowledge the support of the Alfred P. Sloan Foundation through the Deep Carbon Observatory. The basis of this paper was provided by discussion with Deep Carbon Observatory colleagues at the Earth in Five Reactions workshop in Washington, D.C., in March 2018.

References cited

Ague, J.J., and Nicolescu, S. (2014) Carbon dioxide released from subduction zones by fluid-mediated reactions. Nature Geoscience, 7(5), 355–360.10.1038/ngeo2143Search in Google Scholar

Allard, P., Carbonnelle, J., Dajlevic, D., Le Bronec, J., Morel, P., Robe, M., Maurenas, J., Faivre-Pierret, R., Martin, D., and Sabroux, J. (1991) Eruptive and diffuse emissions of CO2 from Mount Etna. Nature, 351, 387–391.10.1038/351387a0Search in Google Scholar

Archer, D., Eby, M., Brovkin, V., Ridgwell, A., Cao, L., Mikolajewicz, U., Caldeira, K., Matsumoto, K., Munhoven, G., and Montenegro, A. (2009) Atmospheric lifetime of fossil fuel carbon dioxide. Annual Review of Earth and Planetary Sciences, 37, 117–134.10.1146/annurev.earth.031208.100206Search in Google Scholar

Behrens, H., Tamic, N., and Holtz, F. (2004) Determination of the molar absorption coefficient for the infrared absorption band of CO2 in rhyolitic glasses. American Mineralogist, 89, 301–306.10.2138/am-2004-2-307Search in Google Scholar

Berner, R.A. (1994) GEOCARB II: A revised model of atmospheric CO2 over phanerozoic time. American Journal of Science, 294(1).10.2475/ajs.294.1.56Search in Google Scholar

Berner, R.A. (1999) A new look at the long-term carbon cycle. GSA Today, 9(11), 1–6.Search in Google Scholar

Berner, R.A., and Caldeira, K. (1997) The need for mass balance and feedback in the geochemical carbon cycle. Geology, 25(10), 955–956.10.1130/0091-7613(1997)025<0955:TNFMBA>2.3.CO;2Search in Google Scholar

Bethke, C.M., Farrell, B., and Yeakel, S. (2018) The Geochemist’s Workbench Release 12.0—Reaction Modeling Guide. Aqueous Solutions, Champaign, Illinois.Search in Google Scholar

Black, B.A., and Gibson, S.A. (2019) Deep carbon and the life cycle of large igneous provinces. Elements, DOI: 10.2138/gselements.15.5.319.10.2138/gselements.15.5.319Search in Google Scholar

Black, B.A., Elkins-Tanton, L.T., Rowe, M.C., and Peate, I.U. (2012) Magnitude and consequences of volatile release from the Siberian Traps. Earth and Planetary Science Letters, 317, 363–373.10.1016/j.epsl.2011.12.001Search in Google Scholar

Black, B.A., Neely, R.R., Lamarque, J.-F., Elkins-Tanton, L.T., Kiehl, J.T., Shields, C.A., Mills, M.J., and Bardeen, C. (2018) Systemic swings in end-Permian climate from Siberian Traps carbon and sulfur outgassing. Nature Geoscience, 11(12), 949.10.1038/s41561-018-0261-ySearch in Google Scholar

Blank, J.G., and Brooker, R.A. (1994) Experimental studies of carbon dioxide in silicate melts; solubility, speciation, and stable carbon isotope behavior. Reviews in Mineralogy and Geochemistry, 30, 157–186.10.1515/9781501509674-011Search in Google Scholar

Blundy, J., Cashman, K.V., Rust, A., and Witham, F. (2010) A case for CO2-rich arc magmas. Earth and Planetary Science Letters, 290(3), 289–301.10.1016/j.epsl.2009.12.013Search in Google Scholar

Boulard, E., Pan, D., Galli, G., Liu, Z., and Mao, W.L. (2015) Tetrahedrally coordinated carbonates in Earth’s lower mantle. Nature Communications, 6, 6311.10.1038/ncomms7311Search in Google Scholar PubMed

Bowen, G.J., Maibauer, B.J., Kraus, M.J., Röhl, U., Westerhold, T., Steimke, A., Gingerich, P.D., Wing, S.L., and Clyde, W.C. (2015) Two massive, rapid releases of carbon during the onset of the Palaeocene–Eocene thermal maximum. Nature Geoscience, 8(1), 44.10.1038/ngeo2316Search in Google Scholar

Brune, S., Williams, S.E., and Müller, R.D. (2017) Potential links between continental rifting, CO2 degassing and climate change through time. Nature Geoscience, 10(12), 941.10.1038/s41561-017-0003-6Search in Google Scholar

Burton, M.R., Sawyer, G.M., and Granieri, D. (2013) Deep carbon emissions from volcanoes. Reviews of Mineralogy and Geochemistry, 75(1), 323–354.10.2138/rmg.2013.75.11Search in Google Scholar

Caldeira, K., and Wickett, M.E. (2003) Oceanography: anthropogenic carbon and ocean pH. Nature, 425, 365.10.1038/425365aSearch in Google Scholar PubMed

Candela, P.A. (1997) A review of shallow, ore-related granites: textures, volatiles, and ore metals. Journal of Petrology, 38(12), 1619–1633.10.1093/petroj/38.12.1619Search in Google Scholar

Caricchi, L., Sheldrake, T.E., and Blundy, J. (2018) Modulation of magmatic processes by CO2 flushing. Earth and Planetary Science Letters, 491, 160–171.10.1016/j.epsl.2018.03.042Search in Google Scholar

Carter, L.B., and Dasgupta, R. (2015) Hydrous basalt–limestone interaction at crustal conditions: Implications for generation of ultracalcic melts and outflux of CO2 at volcanic arcs. Earth and Planetary Science Letters, 427, 202–214.10.1016/j.epsl.2015.06.053Search in Google Scholar

Cashman, K.V. (2004) Volatile controls on magma ascent and eruption. The State of the Planet: Frontiers and Challenges in Geophysics, 109–124.10.1029/150GM10Search in Google Scholar

Cerling, T.E. (1979) Paleochemistry of plio-pleistocene lake Turkana, Kenya. Palaeo-geography, Palaeoclimatology, Palaeoecology, 27, 247–285.10.1016/0031-0182(79)90105-6Search in Google Scholar

Cerling, T. (1994) Chemistry of closed basin lake waters: a comparison between African Rift Valley and some central North American rivers and lakes. The Global Geological Record of Lake Basins, 1, 29–30.Search in Google Scholar

Cerantola, V., Wilke, M., Kantor, I., Ismailova, L., Kupenko, I., McCammon, C., Pascarelli, S., and Dubrovinsky, L.S. (2019) Experimental investigation of FeCO3 (siderite) stability in Earth’s lower mantle using XANES spectroscopy. American Mineralogist, 104, 1083–1091.10.2138/am-2019-6428Search in Google Scholar

Chiodini, G., Frondini, F., and Ponziani, F. (1995) Deep structures and carbon dioxide degassing in central Italy. Geothermics, 24(1), 81–94.10.1016/0375-6505(94)00023-6Search in Google Scholar

Chiodini, G., Frondini, F., Cardellini, C., Parello, F., and Peruzzi, L. (2000) Rate of diffuse carbon dioxide Earth degassing estimated from carbon balance of regional aquifers: the case of central Apennine, Italy. Journal of Geophysical Research: Solid Earth, 105(B4), 8423–8434.10.1029/1999JB900355Search in Google Scholar

Chiodini, G., Cardellini, C., Amato, A., Boschi, E., Caliro, S., Frondini, F., and Ventura, G. (2004) Carbon dioxide Earth degassing and seismogenesis in central and southern Italy. Geophysical Research Letters, 31(7).10.1029/2004GL019480Search in Google Scholar

Chiodini, G., Granieri, D., Avino, R., Caliro, S., Costa, A., Minopoli, C., and Vilardo, G. (2010) Non-volcanic CO2 Earth degassing: Case of Mefite d’Ansanto (southern Apennines), Italy. Geophysical Research Letters, 37(11).10.1029/2010GL042858Search in Google Scholar

Dasgupta, R. (2013) Ingassing, storage, and outgassing of terrestrial carbon through geologic time. Reviews in Mineralogy and Geochemistry, 75, 183–229.10.1515/9781501508318-009Search in Google Scholar

Dasgupta, R., and Hirschmann, M.M. (2010a) The deep carbon cycle and melting in Earth’s interior. Earth and Planetary Science Letters, 298(1-2), 1–13.10.1016/j.epsl.2010.06.039Search in Google Scholar

Dasgupta, R., and Hirschmann, M.M. (2010b) The deep carbon cycle and melting in Earth’s interior. Earth and Planetary Science Letters, 298(1), 1–13.10.1016/j.epsl.2010.06.039Search in Google Scholar

Dasgupta, R., Chi, H., Shimizu, N., Buono, A.S., and Walker, D. (2013) Carbon solution and partitioning between metallic and silicate melts in a shallow magma ocean: Implications for the origin and distribution of terrestrial carbon. Geochimica et Cosmochimica Acta, 102, 191–212.10.1016/j.gca.2012.10.011Search in Google Scholar

Deffeyes, K. S. (1965) Carbonate equilibria: A graphic and algebraic approach 1. Limnology and Oceanography, 10(3), 412–426.10.4319/lo.1965.10.3.0412Search in Google Scholar

DePaolo, D.J. (2015) Sustainable carbon emissions: The geologic perspective. MRS Energy & Sustainability, 2.10.1557/mre.2015.10Search in Google Scholar

Dixon, J.E., and Stolper, E.M. (1995) An experimental study of water and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part II: applications to degassing. Journal of Petrology, 36(6), 1633–1646.Search in Google Scholar

Dixon, J.E., Stolper, E.M., and Holloway, J.R. (1995) An experimental study of water and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part I: calibration and solubility models. Journal of Petrology, 36(6), 1607–1631.Search in Google Scholar

Dixon, J.E., Clague, D.A., Wallace, P., and Poreda, R. (1997) Volatiles in alkalic basalts form the North Arch Volcanic Field, Hawaii: extensive degassing of deep submarineerupted alkalic series lavas. Journal of Petrology, 38(7), 911–939.10.1093/petroj/38.7.911Search in Google Scholar

Doney, S.C., Fabry, V.J., Feely, R.A., and Kleypas, J.A. (2009) Ocean acidification: the other CO2 problem.10.1146/annurev.marine.010908.163834Search in Google Scholar

Dutkiewicz, A., Müller, R.D., Cannon, J., Vaughan, S., and Zahirovic, S. (2018) Sequestration and subduction of deep-sea carbonate in the global ocean since the Early Cretaceous. Geology, 47(1), 91–94.10.1130/G45424.1Search in Google Scholar

Evans, M.J., Derry, L.A., and France-Lanord, C. (2008) Degassing of metamorphic carbon dioxide from the Nepal Himalaya. Geochemistry, Geophysics, Geosystems, 9(4).10.1029/2007GC001796Search in Google Scholar

Facq, S., Daniel, I., Montagnac, G., Cardon, H., and Sverjensky, D.A. (2014) In situ Raman study and thermodynamic model of aqueous carbonate speciation in equilibrium with aragonite under subduction zone conditions. Geochimica et Cosmochimica Acta, 132, 375–390.10.1016/j.gca.2014.01.030Search in Google Scholar

Farrar, C., Sorey, M., Evans, W., Howle, J., Kerr, B., Kennedy, B.M., King, C.-Y., and Southon, J. (1995) Forest-killing diffuse CO2 emission at Mammoth Mountain as a sign of magmatic unrest. Nature, 376, 675–678.10.1038/376675a0Search in Google Scholar

Fine, G., and Stolper, E. (1985) The speciation of carbon dioxide in sodium aluminosilicate glasses. Contributions to Mineralogy and Petrology, 91(2), 105–121.10.1007/BF00377759Search in Google Scholar

Fine, G., and Stolper, E. (1986) Dissolved carbon dioxide in basaltic glasses: concentrations and speciation. Earth and Planetary Science Letters, 76(3), 263–278.10.1016/0012-821X(86)90078-6Search in Google Scholar

Fogel, R.A., and Rutherford, M.J. (1990) The solubility of carbon dioxide in rhyolitic melts; a quantitative FTIR study. American Mineralogist, 75, 1311–1326.Search in Google Scholar

Foley, S.F., and Fischer, T.P. (2017) An essential role for continental rifts and lithosphere in the deep carbon cycle. Nature Geoscience, 10(12), 897.10.1038/s41561-017-0002-7Search in Google Scholar

Friedlingstein, P., Houghton, R., Marland, G., Hackler, J., Boden, T., Conway, T., Canadell, J., Raupach, M., Ciais, P., and Le Quéré, C. (2010) Update on CO2 emissions. Nature Geoscience, 3, 811–812.10.1038/ngeo1022Search in Google Scholar

Garrels, R.M. (1983) The carbonate-silicate geochemical cycle and its effect on atmospheric carbon dioxide over the past 100 million years. American Journal of Science, 283, 641–683.10.2475/ajs.283.7.641Search in Google Scholar

Garrels, R., and MacKenzie, F. (1967) Origin of the chemical compositions of some springs and lakes: equilibrium concepts in natural waters systems. American Chemical Society Advances in Chemistry, 222–242.10.1021/ba-1967-0067.ch010Search in Google Scholar

Gerlach, T. (2011) Volcanic versus anthropogenic carbon dioxide. Eos, Transactions, American Geophysical Union, 92(24), 201–202.10.1029/2011EO240001Search in Google Scholar

Ghiorso, M.S., and Gualda, G.A. (2015) An H2O–CO2 mixed fluid saturation model compatible with rhyolite-MELTS. Contributions to Mineralogy and Petrology, 169(6), 1–30.10.1007/s00410-015-1141-8Search in Google Scholar

Gorman, P.J., Kerrick, D., and Connolly, J. (2006) Modeling open system metamorphic decarbonation of subducting slabs. Geochemistry, Geophysics, Geosystems, 7(4).10.1029/2005GC001125Search in Google Scholar

Guillot, B., and Sator, N. (2011) Carbon dioxide in silicate melts: A molecular dynamics simulation study. Geochimica et Cosmochimica Acta, 75(7), 1829–1857.10.1016/j.gca.2011.01.004Search in Google Scholar

Hauri, E.H., Maclennan, J., McKenzie, D., Gronvold, K., Oskarsson, N., and Shimizu, N. (2018) CO2 content beneath northern Iceland and the variability of mantle carbon. Geology, 46(1), 55–58.10.1130/G39413.1Search in Google Scholar

Hauri, E.H., Cottrell, E., Kelley, K.A., Tucker, J.M., Shimizu, K., Le Voyer, M., Marske, J., and Saal, A.E. (2019) Carbon in the Convecting Mantle. Earth’s Deep Carbon: past to present. Cambridge.10.1017/9781108677950.009Search in Google Scholar

Hirschmann, M.M. (2018) Comparative deep Earth volatile cycles: The case for C recycling from exosphere/mantle fractionation of major (H2O, C, N) volatiles and from H2O/Ce, CO2/Ba, and CO2/Nb exosphere ratios. Earth and Planetary Science Letters, 502, 262–273.10.1016/j.epsl.2018.08.023Search in Google Scholar

Hirschmann, M.M., and Dasgupta, R. (2009) The H/C ratios of Earth’s near-surface and deep reservoirs, and consequences for deep Earth volatile cycles. Chemical Geology, 262(1-2), 4–16.10.1016/j.chemgeo.2009.02.008Search in Google Scholar

Hunt, J.A., Zafu, A., Mather, T.A., Pyle, D.M., and Barry, P.H. (2017) Spatially variable CO2 degassing in the Main Ethiopian Rift: Implications for magma storage, volatile transport, and rift-related emissions. Geochemistry, Geophysics, Geosystems, 18(10), 3714–3737.10.1002/2017GC006975Search in Google Scholar

Hutchison, W., Mather, T.A., Pyle, D.M., Biggs, J., and Yirgu, G. (2015) Structural controls on fluid pathways in an active rift system: A case study of the Aluto volcanic complex. Geosphere, 11(3), 542–562.10.1130/GES01119.1Search in Google Scholar

Kelemen, P.B., and Manning, C.E. (2015) Reevaluating carbon fluxes in subduction zones, what goes down, mostly comes up. Proceedings of the National Academy of Sciences, 112(30), E3997–E4006.10.1073/pnas.1507889112Search in Google Scholar

Kerrick, D.M., and Caldeira, K. (1998) Metamorphic CO2 degassing from orogenic belts. Chemical Geology, 145(3-4), 213–232.10.1016/S0009-2541(97)00144-7Search in Google Scholar

Kerrick, D., and Connolly, J. (2001) Metamorphic devolatilization of subducted marine sediments and the transport of volatiles into the Earth’s mantle. Nature, 411, 293–296.10.1038/35077056Search in Google Scholar PubMed PubMed Central

Konschak, A., and Keppler, H. (2014) The speciation of carbon dioxide in silicate melts. Contributions to Mineralogy and Petrology, 167(5), 998.10.1007/s00410-014-0998-2Search in Google Scholar

Lee, C.-T.A., Shen, B., Slotnick, B.S., Liao, K., Dickens, G.R., Yokoyama, Y., Lenardic, A., Dasgupta, R., Jellinek, M., and Lackey, J.S. (2013) Continental arc–island arc fluctuations, growth of crustal carbonates, and long-term climate change. Geosphere, 9(1), 21–36.10.1130/GES00822.1Search in Google Scholar

Lee, H., Muirhead, J.D., Fischer, T.P., Ebinger, C.J., Kattenhorn, S.A., Sharp, Z.D., and Kianji, G. (2016) Massive and prolonged deep carbon emissions associated with continental rifting. Nature Geoscience, 9(2), 145.10.1038/ngeo2622Search in Google Scholar

Lee, C.T., Jiang, H., Dasgupta, R., and Torres, M. (2019) A Framework for Understanding Whole-Earth Carbon Cycling. In B. Orcutt, I. Daniel, and R. Dasgupta, Eds., Earth’s Deep Carbon: Past to Present, pp. 313–357. Cambridge.10.1017/9781108677950.011Search in Google Scholar

Lenton, T.M., and Britton, C. (2006) Enhanced carbonate and silicate weathering accelerates recovery from fossil fuel CO2 perturbations. Global Biogeochemical Cycles, 20(3).10.1029/2005GB002678Search in Google Scholar

Li, J., and Fei, Y. (2003) Experimental constraints on core composition. Treatise on Geochemistry, 2, 568.Search in Google Scholar

Li, J., Chen, B., Mookherjee, M., and Morard, G. (2019) Carbon versus other light elements in Earth’s core. In B. Orcutt, I. Daniel, and R. Dasgupta, Eds., Earth’s Deep Carbon: Past to Present, p. 40–65. Cambridge.10.1017/9781108677950.003Search in Google Scholar

Lilley, M.D., Butterfield, D.A., Lupton, J.E., and Olson, E.J. (2003) Magmatic events can produce rapid changes in hydrothermal vent chemistry. Nature, 422, 878.10.1038/nature01569Search in Google Scholar PubMed

Lund, D.C., Asimow, P.D., Farley, K.A., Rooney, T.O., Seeley, E., Jackson, E.W., and Durham, Z.M. (2016) Enhanced East Pacific Rise hydrothermal activity during the last two glacial terminations. Science, 351(6272), 478–482.10.1126/science.aad4296Search in Google Scholar PubMed

Lupton, J., Lilley, M., Butterfield, D., Evans, L., Embley, R., Massoth, G., Christenson, B., Nakamura, K.i., and Schmidt, M. (2008) Venting of a separate CO2-rich gas phase from submarine arc volcanoes: Examples from the Mariana and Tonga-Kermadec arcs. Journal of Geophysical Research: Solid Earth, 113(B8).10.1029/2007JB005467Search in Google Scholar

Mason, E., Edmonds, M., and Turchyn, A.V. (2017) Remobilization of crustal carbon may dominate volcanic arc emissions. Science, 357, 290–294.10.1126/science.aan5049Search in Google Scholar PubMed

McCammon, C., Bureau, H., Cleaves, H.J. II, Cottrell, E., Dorfman, S.M., Kellogg, L.H., Li, J., Mikhail, S., Moussallam, Y., Sanloup, C., and others. (2020) Deep Earth carbon reactions through time and space. American Mineralogist, 105, 22–27.10.2138/am-2020-6888CCBYSearch in Google Scholar

Métrich, N., and Wallace, P.J. (2008) Volatile abundances in basaltic magmas and their degassing paths tracked by melt inclusions. Reviews in Mineralogy and Geochemistry, 69, 363–402.10.1515/9781501508486-011Search in Google Scholar

Moore, G. (2008) Interpreting H2O and CO2 contents in melt inclusions: constraints from solubility experiments and modeling. Reviews in Mineralogy and Geochemistry, 69, 333–362.10.1515/9781501508486-010Search in Google Scholar

Moore, G., Vennemann, T., and Carmichael, I.S.E. (1998) An empirical model for the solubility of H2O in magmas to 3 kilobars. American Mineralogist, 83, 36–42.10.2138/am-1998-1-203Search in Google Scholar

Morizet, Y., Kohn, S., and Brooker, R. (2001) Annealing experiments on CO2-bearing jadeite glass: an insight into the true temperature dependence of CO2 speciation in silicate melts. Mineralogical Magazine, 65(6), 701–707.10.1180/0026461016560001Search in Google Scholar

Morizet, Y., Vuilleumier, R., and Paris, M. (2015) A NMR and molecular dynamics study of CO2-bearing basaltic melts and glasses. Chemical Geology, 418, 89–103.10.1016/j.chemgeo.2015.03.021Search in Google Scholar

Mörner, N.-A., and Etiope, G. (2002) Carbon degassing from the lithosphere. Global and Planetary Change, 33(1), 185–203.10.1016/S0921-8181(02)00070-XSearch in Google Scholar

Moussallam, Y., Morizet, Y., Massuyeau, M., Laumonier, M., and Gaillard, F. (2015) CO2 solubility in kimberlite melts. Chemical Geology, 418, 198–205.10.1016/j.chemgeo.2014.11.017Search in Google Scholar

Moussallam, Y., Florian, P., Corradini, D., Morizet, Y., Sator, N., Vuilleumier, R., Guillot, B., Iacono-Marziano, G., Schmidt, B.C., and Gaillard, F. (2016) The molecular structure of melts along the carbonatite–kimberlite–basalt compositional joint: CO2 and polymerisation. Earth and Planetary Science Letters, 434, 129–140.10.1016/j.epsl.2015.11.025Search in Google Scholar

Mysen, B.O., Arculus, R., and Eggler, D.H. (1975) Solubility of carbon dioxide in melts of andesite, tholeiite, and olivine nephelinite composition to 30 kbar pressure. Contributions to Mineralogy and Petrology, 53(4), 227–239.10.1007/BF00382441Search in Google Scholar

Newman, S., and Lowenstern, J.B. (2002) VolatileCalc: a silicate melt–H2O–CO2 solution model written in Visual Basic for excel. Computers & Geosciences, 28(5), 597–604.10.1016/S0098-3004(01)00081-4Search in Google Scholar

Nowak, M., Porbatzki, D., Spickenbom, K., and Diedrich, O. (2003) Carbon dioxide speciation in silicate melts: a restart. Earth and Planetary Science Letters, 207, 131–139.10.1016/S0012-821X(02)01145-7Search in Google Scholar

Ogden, D.E., and Sleep, N.H. (2012) Explosive eruption of coal and basalt and the end-Permian mass extinction. Proceedings of the National Academy of Sciences, 109(1), 59–62.10.1073/pnas.1118675109Search in Google Scholar PubMed PubMed Central

Orr, J.C., Fabry, V.J., Aumont, O., Bopp, L., Doney, S.C., Feely, R.A., Gnanadesikan, A., Gruber, N., Ishida, A., and Joos, F. (2005) Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature, 437, 681.10.1038/nature04095Search in Google Scholar PubMed

Pan, D., and Galli, G. (2016) The fate of carbon dioxide in water-rich fluids under extreme conditions. Science Advances, 2(10), e1601278.10.1126/sciadv.1601278Search in Google Scholar PubMed PubMed Central

Papale, P., Moretti, R., and Barbato, D. (2006) The compositional dependence of the saturation surface of H2O+CO2 fluids in silicate melts. Chemical Geology, 229(1-3), 78–95.10.1016/j.chemgeo.2006.01.013Search in Google Scholar

Reichow, M.K., Pringle, M., Al’Mukhamedov, A., Allen, M., Andreichev, V., Buslov, M., Davies, C., Fedoseev, G., Fitton, J., and Inger, S. (2009) The timing and extent of the eruption of the Siberian Traps large igneous province: Implications for the end-Permian environmental crisis. Earth and Planetary Science Letters, 277(1), 9–20.10.1016/j.epsl.2008.09.030Search in Google Scholar

Saal, A.E., Hauri, E.H., Langmuir, C.H., and Perfit, M.R. (2002) Vapour undersaturation in primitive mid-ocean-ridge basalt and the volatile content of Earth’s upper mantle. Nature, 419, 451–455.10.1038/nature01073Search in Google Scholar PubMed

Sabine, C.L., Feely, R.A., Gruber, N., Key, R.M., Lee, K., Bullister, J.L., Wanninkhof, R., Wong, C., Wallace, D.W., and Tilbrook, B. (2004) The oceanic sink for anthropogenic CO2 Science, 305, 367–371.10.1126/science.1097403Search in Google Scholar PubMed

Saunders, A.D. (2005) Large igneous provinces: origin and environmental consequences. Elements, 1(5), 259–263.10.2113/gselements.1.5.259Search in Google Scholar

Shirey, S.B., Cartigny, P., Frost, D.J., Keshav, S., Nestola, F., Nimis, P., Pearson, D.G., Sobolev, N.V., and Walter, M.J. (2013) Diamonds and the geology of mantle carbon. Reviews in Mineralogy and Geochemistry, 75, 355–421.10.2138/rmg.2013.75.12Search in Google Scholar

Shishkina, T.A., Botcharnikov, R.E., Holtz, F., Almeev, R.R., Jazwa, A.M., and Jakubiak, A.A. (2014) Compositional and pressure effects on the solubility of H2O and CO2 in mafic melts. Chemical Geology, 388, 112–129.10.1016/j.chemgeo.2014.09.001Search in Google Scholar

Sleep, N.H., and Zahnle, K. (2001) Carbon dioxide cycling and implications for climate on ancient Earth. Journal of Geophysical Research: Planets, 106(E1), 1373–1399.10.1029/2000JE001247Search in Google Scholar

Stewart, E.M., Ague, J.J., Ferry, J.M., Schriffries, C.M., Tao, R.-B., Isson, T.T., and Planavsky, N.J. (2019) Carbonation and decarbonation reactions: Implications for planetary habitability. American Mineralogist, 104, 1369–1380.10.2138/am-2019-6884Search in Google Scholar

Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V., and Midgley, P.M. (2013) Climate change 2013: The physical science basis. Intergovernmental Panel on Climate Change, Working Group I Contribution to the IPCC Fifth Assessment Report (AR5), 25. Cambridge.Search in Google Scholar

Stolper, E., and Holloway, J.R. (1988) Experimental determination of the solubility of carbon dioxide in molten basalt at low pressure. Earth and Planetary Science Letters, 87(4), 397–408.10.1016/0012-821X(88)90004-0Search in Google Scholar

Stumm, W., and Morgan, J.J. (1996) Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. Wiley.Search in Google Scholar

Sundquist, E.T. (1993) The global carbon dioxide budget. Science, 934–941.10.1126/science.259.5097.934Search in Google Scholar

Svensen, H., Planke, S., Polozov, A.G., Schmidbauer, N., Corfu, F., Podladchikov, Y. Y., and Jamtveit, B. (2009) Siberian gas venting and the end-Permian environmental crisis. Earth and Planetary Science Letters, 277(3), 490–500.10.1016/j.epsl.2008.11.015Search in Google Scholar

Sverjensky, D.A., and Huang, F. (2015) Diamond formation due to a pH drop during fluid–rock interactions. Nature Communications, 6, 8702.10.1038/ncomms9702Search in Google Scholar PubMed PubMed Central

Tamburello, G., Pondrelli, S., Chiodini, G., and Rouwet, D. (2018) Global-scale control of extensional tectonics on CO2 Earth degassing. Nature Communications, 9(1), 4608.10.1038/s41467-018-07087-zSearch in Google Scholar PubMed PubMed Central

Tutolo, B.M., and Tosca, N.J. (2018) Experimental examination of the Mg-silicatecarbonate system at ambient temperature: Implications for alkaline chemical sedimentation and lacustrine carbonate formation. Geochimica et Cosmochimica Acta, 225, 80–101.10.1016/j.gca.2018.01.019Search in Google Scholar

Wallace, P.J., Anderson, A.T., and Davis, A.M. (1999) Gradients in H2O, CO2 and exsolved gas in a large-volume silicic magma system: Interpreting the record preserved in melt inclusions from the Bishop Tuff. Journal of Geophysical Research: Solid Earth, 104(B9), 20097–20122.10.1029/1999JB900207Search in Google Scholar

Werner, C., and Brantley, S. (2003) CO2 emissions from the Yellowstone volcanic system. Geochemistry, Geophysics, Geosystems, 4, 1061, doi:10.1029/2002GC000473.10.1029/2002GC000473Search in Google Scholar

Werner, C., Fischer, T.P., Aiuppa, A., Edmonds, M., Cardellini, C., Carn, S., Chiodini, G., Cottrell, E., Burton, M., Shinohara, H., and Allard, P. (2019) Carbon dioxide emissions from subaerial volcanic regions: Two decades in review. In B. Orcutt, I. Daniel, and R. Dasgupta, Eds., Earth’s Deep Carbon: Past to Present, pp. 188–236. Cambridge University Press.10.1017/9781108677950.008Search in Google Scholar

Winguth, A.M., Thomas, E., and Winguth, C. (2012) Global decline in ocean ventilation, oxygenation, and productivity during the Paleocene-Eocene Thermal Maximum: Implications for the benthic extinction. Geology, 40(3), 263–266.10.1130/G32529.1Search in Google Scholar

Wood, B.J. (1993) Carbon in the core. Earth and Planetary Science Letters, 117(3-4), 593–607.10.1016/0012-821X(93)90105-ISearch in Google Scholar

Wood, B.J., Li, J., and Shahar, A. (2013) Carbon in the core: its influence on the properties of core and mantle. Reviews in Mineralogy and Geochemistry, 75, 231–250.10.1515/9781501508318-010Search in Google Scholar

Wright, V.P. (2012) Lacustrine carbonates in rift settings: the interaction of volcanic and microbial processes on carbonate deposition. Geological Society, London, Special Publications, 370, SP370, 2.10.1144/SP370.2Search in Google Scholar

Zachos, J.C., Röhl, U., Schellenberg, S.A., Sluijs, A., Hodell, D.A., Kelly, D.C., Thomas, E., Nicolo, M., Raffi, I., and Lourens, L.J. (2005) Rapid acidification of the ocean during the Paleocene-Eocene thermal maximum. Science, 308, 1611–1615.10.1126/science.1109004Search in Google Scholar PubMed

Zeebe, R.E., Zachos, J.C., Caldeira, K., and Tyrrell, T. (2008) Carbon emissions and acidification. Science, 321, 51–52.10.1126/science.1159124Search in Google Scholar PubMed

Zeebe, R.E., Ridgwell, A., and Zachos, J.C. (2016) Anthropogenic carbon release rate unprecedented during the past 66 million years. Nature Geoscience, 9(4), 325.10.1038/ngeo2681Search in Google Scholar

Received: 2019-02-01
Accepted: 2019-08-20
Published Online: 2019-12-30
Published in Print: 2020-01-28

© 2020 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.2138/am-2020-6986CCBY/html
Scroll to top button