Introduction

Modern open-ocean sediments are dominated by phytoplankton. Calcium carbonate, chiefly in sediments deposited above the calcite compensation depth, is essentially calcite forming the shells of organisms such as coccolithophores. It displays carbon isotopic ratios comparable to bicarbonate ions dissolved in seawater, characterized by δ13C ≈ 0‰ expressed as 13C/12C per mil difference normalized to the international Vienna-Pee Dee Belemnite (VPDB) standard. Conversely organic matter, which is essentially phytoplankton debris in the open seafloor, is depleted at 13C due to fractionation effects induced by photosynthesis, such that δ13C ≈ –20‰ VPDB1. Stable carbon isotopes and mass-balance calculations relying on simple mixing models have been used extensively to determine the relative contribution of organic matter and marine carbonates in sedimentary rocks2, in their metamorphic equivalents3 and in volcanic arc emissions4. However, the applicability of simple mixing models hinges on assumptions that may be overly simplistic, including that the sedimentary “end-member” compositions are spatially and temporally invariant, and that isotopic equilibrium is attained during metamorphism. The latter is particularly problematic because sedimentary organic carbon and coexisting carbonate may exhibit substantial isotopic exchange with increasing metamorphic grade, in particular at temperatures >650 °C during prograde graphitization3,5,6,7. In turn, fully crystalline graphite displays a sluggish rate of isotopic diffusion and may be unaffected by isotopic reset even in cases of intense metamorphism and fluid/rock interactions8.

Carbonates and graphitic carbon derived from organic matter, formerly assumed to be refractory to dissolution and devolatilization during subduction9,10, are now thought to show a non-negligible solubility in subduction fluids at least at certain PTfO2–pH conditions11,12,13,14,15,16,17,18. The interaction of subducted sediments with deep fluids14,19 produces dissolved carbon that is transferred from the slab to the overlying mantle wedge, prompting carbonation/metasomatism and/or partial melting20,21,22,23, and eventually returning to the surface via CO2 emitted by arc volcanoes4. The global arc average δ13C is −2.8 to −3.3 ‰4. This is heavier than the major carbon isotopic composition signature of the upper mantle (δ13C = −6.0 ± 2.5‰)4, but significantly lighter than sedimentary carbonates (δ13C ≈ 0‰24). As it is generally assumed that the carbon isotope signature of arc emissions reflects that of their source, relatively high δ13C values would point out assimilation of shallow “crustal” limestones, while low δ13C are usually attributed to subducted organic carbon4. However, processes of dissolution and of isotopic exchange involving organic and inorganic carbon beneath arcs are still not fully understood.

In this work, we investigate the carbon isotopic exchange occurring in a model system representative of open-ocean sediments containing calcium carbonate + organic matter subducted at subarc depths and interacting with aqueous fluids rising from the underlying dehydrating oceanic lithosphere25. We provide the quantitative chemical analysis of the volatile species and the measured carbon isotopic composition of CO2 produced by dissolution in water of graphite and of aragonite at P = 3 GPa, T = 700 °C and at redox-controlled conditions buffered to fH2 = FMQ (equivalent to fO2 expressed as ∆FMQ = + 0.61 log units). These conditions are selected on the basis of the predicted peak of CO2 produced by oxidative dissolution of graphite in subduction zones (Supplementary Fig. 1). Starting materials of synthetic labelled CaCO3 (99.4% 13C) and synthetic graphite (98.9% 12C) are used as analogues for natural “heavy” carbonate and “light” organic matter and to generate a maximum isotopic difference in experiments (close to pure 13C and 12C end-members). Control experiments are performed with oxalic acid di-hydrate (98.7% 12C) as the source of CO2 instead of graphite. The dissolution process and the effect of different CO2/aragonite ratios and of different run durations from 0.24 to 240 h are evaluated and compared with thermodynamic calculations that include consideration of aqueous solute speciation26,27,28 or consider only gas speciation in a conventional graphite-saturated COH fluid model29,30. Finally, a comparison with conventional isotopic models is provided. Experimental data allow the development of mathematical models extending the applicability of the results to a wide range of redox conditions and of fluid/rock ratios in order to predict the δ13C of CO2 released from the sedimentary slab and to envisage conditions required to meet the global average arc signature.

Results and discussion

CO2 evolved by the aqueous dissolution of aragonite-only, graphite-only and mixed aragonite + graphite

In all experiments we observed that aragonite crystals in run products displayed textural evidence of dissolution-reprecipitation, including step edges, and fine-grained recrystallized rims around larger relict cores (Fig. 1). Conversely, we never observed texturally precipitation of newly formed graphite or graphite recrystallization, nor evidence for graphite isotopic variation compared to its starting composition.

Fig. 1: Electron microscope images showing microtextures of run products.
figure 1

a Recrystallized aragonite, showing rims of sub-micrometric crystals surrounding relict cores 50–100 µm in size. b Assemblage graphite + aragonite; aragonite crystals show dissolution/reprecipitation microtextures such as crystal size reduction, hoppering, step edges and euhedral crystal intergrowth.

In aragonite-only runs, despite the evidence of dissolution–precipitation microtextures, the absolute amount of CO2 evolved by the aqueous dissolution of aragonite-only is very low, close to the analytical detection limit, with XCO2 [=CO2(aq)/(H2O + CO2(aq))molar] = 0.001 corresponding to 0.041 mol CO2/kg H2O (Fig. 2a; Supplementary Tables 1 and 2). We compared this result with thermodynamic modelling performed at the investigated PTfO2 conditions using the Deep Earth Water (DEW) model26,28 (Fig. 2b; Supplementary Table 3). Calculations indicate that fluids in equilibrium with pure aragonite should display a basic pH = 5.09 (neutral pH = 3.09 at 3 GPa and 700 °C) and XCO2 = 0.0002, which is even lower than the measured value. However, the model predicts that the dominant carbon-bearing dissolution product of aragonite at the investigated conditions is the calcium-bicarbonate ion Ca(HCO3)+, in agreement with previous studies13,14. According to the model, a substantial concentration of carbon is present in the form of ionic species. In particular, Ca(HCO3)+ accounts for the 74.3 mol% of the total carbon-bearing dissolved species, while CO2(aq) only 7.3 mol% (Supplementary Table 3). The solubility of aragonite is therefore higher than that inferred on the basis of the measured CO2 only. As Ca(HCO3)+ is not measurable by QMS, we rely on the predicted Ca(HCO3)+ abundance to correct up the bulk aragonite solubility to 0.459 mol CO2/kg H2O, equivalent to 5.28 × 103 ppm (=mg C/kg solution; 495 ppm of which deriving from measured CO2(aq)), which agrees well with previous estimates14. In aragonite-only runs, the evolved CO2 is independent of fO2, as demonstrated by the nearly identical XCO2 = 0.002 in the more oxidized (≈ΔFMQ + 2) control run buffered by Re–ReO2 (RRO in Fig. 2a; Supplementary Tables 1 and 2). On the contrary, the addition of ~0.5 molal to ~1 molal chlorine to lower pH in control experiments effectively boosts fluid CO2 concentrations to 2.17–2.46 mol CO2/kg H2O (corresponding to XCO2 = 0.038–0.042; Supplementary Tables 1 and 2).

Fig. 2: Experimental CO2 contents and thermodynamic modelling of fluids.
figure 2

a CO2 concentration (molality) in experimental aqueous fluids interacting with (i) aragonite-only (blue), (ii) graphite-only (grey) and (iii) graphite + aragonite (magenta). Typical analytical uncertainty is 1 mol%. Dashed line: CO2 content predicted by thermodynamic modelling of graphite-saturated COH fluids. Run duration (h) is shown at the top of each bar. RRO: run buffered by Re–ReO2 (ΔFMQ ≈ +2) instead of ferrosilite + magnetite + coesite; b fO2 vs. pH diagram at 3 GPa and 700 °C generated by thermodynamic modelling (Deep Earth Water model26,28) of aqueous fluids in equilibrium with aragonite-only (blue dot), graphite-only (grey dot) and aragonite + graphite (magenta dot). Black solid lines: saturation curves. Coloured dots: experimental conditions at fO2 buffered by ferrosilite + magnetite + coesite. Grey field: CO2(aq) is the dominant carbon-bearing species; it is adjacent to Ca(HCO3)+-dominated field at higher pH and to CH4(aq)-dominated field at lower fO2 values. Neutral pH is shown for reference with a dashed line. Calculations are performed at 3 GPa, 700 °C and fH2 buffered by ferrosilite + magnetite + coesite + H2O (equivalent to log (fO2 /1 bar)= −13.36; ΔFMQ = +0.61).

Compared to runs with aragonite-only, the CO2 amount in fluids in experiments with graphite-only is significantly higher, with XCO2 = 0.339 after 240 h (Fig. 2a) corresponding to 28.5 mol CO2/kg H2O and a carbon concentration of 1.52 × 105 ppm, two orders of magnitude higher than in fluids dissolving aragonite-only (Fig. 2a; Supplementary Tables 1 and 2). DEW model calculations show that fluids in equilibrium with pure graphite should display an acidic pH = 2.21 (Fig. 2b) and XCO2 = 0.305 (Supplementary Table 3), which is nearly identical to the value of XCO2 = 0.303 predicted by conventional modelling of graphite-saturated COH fluids29,30 (Fig. 2a; Supplementary Table 4) and close to the measured value of 0.339. In fluids in equilibrium with graphite-only, CO2(aq) is predicted to account for 99.93% of the total dissolved carbon (Supplementary Table 3).

The CO2 amount measured in time-resolved runs containing mixed aragonite and graphite (Fig. 2a; Tables 1 and 2) ranges from XCO2 = 0.026 after 0.24 h to 0.166 after 240 h. Results suggest that chemical equilibrium is almost reached after 72 h. After 240 h, fluids contain 11.19 mol CO2/kgwater corresponding to 8.96 × 104 ppm C, about one half that in fluids in equilibrium with graphite-only. This decline is not anticipated by either the DEW model calculations–which predict only a slightly lower XCO2 of 0.284 for fluids in equilibrium with graphite-only (Supplementary Table 3)—or by the graphite-saturated COH fluid model—which cannot account for the dissolution of aragonite because it does not consider components other than C, O and H. Our experimental result confirms that the estimation of XCO2 with available thermodynamic models is hampered in complex systems15. Nevertheless the DEW model is still useful to predict that fluids in equilibrium with both aragonite and graphite display a basic pH = 4.2 (Supplementary Table 3), which is higher than fluids in equilibrium with graphite-only but lower than fluids in equilibrium with aragonite-only, dominated by the dissolution product Ca(HCO3)+ (Fig. 2b). The consequence is that, as in fluids in equilibrium with graphite alone, CO2(aq) is calculated to predominate in fluids in equilibrium with both graphite and aragonite (Fig. 2b), with CO2(aq) accounting for 91.4% of the total carbon-bearing species, whereas Ca(HCO3)+ constitutes only 3.70% of the total carbon in solution (Supplementary Table 3).

Carbon isotopic composition of graphite, aragonite and CO2

In experiments where water interacts with single minerals, the isotopic abundances of evolved CO2 are identical to that of the starting materials within analytical uncertainties. In particular, due to the analytical sensitivity of QMS, this is evident in runs characterized by higher CO2 production, i.e. 13C-aragonite + chlorine and 12C-graphite-only runs. In graphite-alone runs, the average CO2 isotopic abundance is 1.11% 13C after 240 h (Supplementary Table 1), indistinguishable from the 13C abundance of the starting graphite (1.09% 13C measured by IRMS; Supplementary Fig. 2) considering analytical uncertainties.

The isotopic signature of evolved CO2 in experiments with both 13C-aragonite and 12C-graphite (Fig. 3a; Supplementary Table 1) is relatively rich in 12C (13C = 52.9–62.8%) only in runs characterized by a very short duration of 0.24 h, while it becomes relatively homogeneous for t ≥ 24 h reaching 13C abundance = 82.2% after 240 h. Bulk measurements by means of total carbonate HCl dissolution allowed retrieval of the final average isotopic composition of after-run aragonite coexisting with CO2 (y-axis in Fig. 3a; Supplementary Table 1). Low 13C abundances in evolved CO2 are always coupled with high 13C in bulk aragonite, the latter tending to the starting 13C-CaCO3 value of 99.4% 13C (blue dot in Fig. 3a) in the shortest 0.24 h (13C = 96.1–97.2%) and 2.4 h (13C = 95.8%) runs (Fig. 3a; Supplementary Table 1). In longer t ≥ 24 h runs, 13C abundances of CO2 and of bulk carbonate converge (cf. black dashed line in Fig. 3a). After 240 h, the 13C abundance in aragonite is 82.9%, nearly identical to that of evolved CO2 (i.e. 82.2%).

Fig. 3: Isotopic composition of experimental fluids and solids.
figure 3

a Measured 13C abundance (%) in CO2 versus in bulk after-run aragonite. Grey reference line indicates identical 13C abundances in CO2 and in bulk aragonite. Starting 13C% of carbonate and graphite are indicated with blue and grey dots, respectively. Representative point analyses of aragonite cores and rims are shown with black dashed lines. Magenta dots: 13C-CaCO3 + graphite runs. Black stars: 13C-CaCO3 + oxalic acid di-hydrate (OAD; starting 13C% is coincident with starting graphite) runs. b CO2/13C-CaCO3 molar ratio plotted against 13C abundance (%) in CO2 of runs with duration >24 h. Dashed line represents the best fit of the data using an exponential equation. c Conceptual model for the carbon isotope exchange observed in experiments. At step 1, isotopically pure 12C-graphite (brown) and 13C-aragonite (blue) coexist in the same assemblage. As the interaction with water starts (step 2), graphite undergoes oxidative dissolution forming 12CO2(aq), while aragonite dissolves forming Ca(H13CO3)+, as predicted by thermodynamic modelling. While graphite oxidative dissolution is self-limiting because the maximum amount of CO2(aq) in the fluid is constrained by the redox state of the system, aragonite undergoes a continuous process of dissolution/precipitation (step 3). Therefore, the final isotopic composition of CO2(aq), which is in dynamic equilibrium with Ca(HCO3)+ and thus with aragonite, becomes rapidly enriched in 13C ending with 13C abundances of >80% of the starting 13C-aragonite.

Micro-analyses performed by means of NanoSIMS and LA-ICP-MS (dashed lines in Fig. 3a; Supplementary Tables 5 and 6) show that aragonite crystals are isotopically zoned, in particular in short runs, with recrystallized rims characterized by 13C abundances comparable to that of bulk aragonite and of CO2 in longer runs (e.g. 13C = 81.6(8)% in the 24 h run COH106; Supplementary Table 6), and relict cores showing high 13C abundance up to 98(1)% 13C (e.g. run COH97; Supplementary Table 6) approaching the starting composition of 13C-CaCO3. Conversely, the 13C abundance in after-run graphite indicates negligible isotopic re-equilibration (Supplementary Fig. 2; Supplementary Table 5, 6). NanoSIMS measurements range from 1.132(6)% to 1.195(3)% (Supplementary Fig. 2; Supplementary Table 5), concordant with LA-ICP-MS measurements ranging from 13C = 1.1(1)% to 1.5(2)% (Supplementary Fig. 2; Supplementary Table 6).

The 12C-rich composition of CO2 in the shortest runs and the lack of newly formed graphite suggest that in mixed aragonite–graphite runs, the early source of CO2 was graphite undergoing irreversible oxidation. To validate this hypothesis, we performed a series of control experiments at identical experimental conditions but without graphite, where the early source of 12CO2 is provided by oxalic acid di-hydrate (13C = 1.29%; OAD and black stars in Fig. 3a), which decomposes already at T ≈ 200 °C to a mixed H2O–CO2 fluid31. OAD has been added in the proper amount to keep the same CO2/13C-CaCO3 ratio ≈ 0.1 characterizing graphite + aragonite experiments (Fig. 3b; Supplementary Table 1). In addition, 10 times more OAD was added in a single run to increase the CO2/13C-CaCO3 ratio to a value of ~0.5 (“enhanced CO2” in Fig. 3a, b; Supplementary Table 1). In OAD + 13C-CaCO3 runs with CO2/13C-CaCO3 ratio ≈ 0.1, the isotopic composition of both CO2 and bulk carbonate shows no appreciable differences compared to aragonite + graphite runs, further supporting the hypothesis of graphite as the early source of CO2. In addition, the OAD + 13C-CaCO3 run with CO2/13C-CaCO3 ratio ≈ 0.5 makes evident that the 13C abundance of CO2 (and of coexisting aragonite) varies as a function of the CO2/13C-aragonite ratio (CAR), which can be described by the following exponential function (Fig. 3b):

$${}^{13}{{{{{\rm{C}}}}}}_{{{{{{{\rm{CO}}}}}}}_{2}} \% =100\times {e}^{\left(-1.5\times \!{{{{{\rm{CAR}}}}}}\right)}$$
(1)

which represents a mathematical model for the competing isotopic buffering in fluids where CO2 originated from the oxidative dissolution of 12C-graphite interacts with 13C-aragonite. Equation 1 shows that graphite-derived CO2 (and coexisting aragonite) are fully isotopically buffered by 13C-aragonite to 13C abundances ≈ 100% when CAR tends to zero. Conversely, by increasing CAR the contribution of 12C-graphite to the isotopic composition of CO2 becomes increasingly important, with 13C abundances tending to zero (i.e. fully buffered by graphite) at very high CO2/13C-aragonite ratios (e.g. 13CCO2 = 0.01 when CAR = 6).

A conceptual model for carbon isotopic exchange among graphite, aragonite and CO2

In runs containing mixed 13C-aragonite and 12C-graphite, the isotopic composition of CO2 after 24 h is nearly coincident with that of the recrystallized aragonite, although it is produced mainly by the oxidative dissolution of graphite which remains isotopically unchanged. Our experimental results confirm that carbon isotope exchange between graphite and aragonite at 700 °C is sluggish, in agreement with previous findings suggesting that carbon diffusion in graphite is very slow at T < 1300 °C32,33. In the shortest 0.24 h and 2.4 h runs, despite the similar 13C abundances of CO2 and aragonite rims, the bulk carbonate 13C is markedly higher because of a number of unreacted 13C-rich cores persist, as shown by NanoSIMS and LA-ICP-MS analyses. In order to develop a conceptual model for the isotopic exchange among aragonite, graphite and CO2, we rely on observed microtextures, measurements and thermodynamic modelling results suggesting that during the run both graphite and aragonite undergo dissolution. However, 12C-graphite produces mainly 12CO2 by irreversible oxidation, which can be expressed by the following fO2-dependent reaction:

$$^{12}{{\rm{C}}}+{{{{{\rm{O}}}}}}_{2}= > ^{12}{{{{{\rm{CO}}}}}}_{2}$$
(2)

13C-aragonite dissolves initially forming mainly calcium-bicarbonate ions (Ca(H13CO3)+; Fig. 3c), according to the reactions involving the predominant species:

$${{{{{{\rm{Ca}}}}}}{}^{13}{{\rm{C}}}{{{{{\rm{O}}}}}}_{3}+{{{{{\rm{H}}}}}}_{2}{{{{{\rm{O}}}}}} < = > {{{{{\rm{Ca}}}}}}({{{{{\rm{H}}}}}}{}^{13}{{{{{\rm{CO}}}}}}_{3})^{+}+{{{{{\rm{OH}}}}}}}^{-}$$
(3)

when the fluid becomes saturated with respect to graphite, CO2 production by graphite oxidation stops and the unreacted graphite remains chemically and isotopically inert, while aragonite continuously dissolves and reprecipitates in dynamic equilibrium with the dissolved aqueous carbon species CO2 and Ca(HCO3)+, which in turn exchange carbon isotopes according to the following reaction:

$$^{12}{{{{{\rm{CO}}}}}}_{2} +{{{{{\rm{Ca}}}}}}({{{{{\rm{H}}}}}}{}^{13}{{{{{\rm{CO}}}}}}_{3})^{+} < = > {}^{13}{{{{{\rm{CO}}}}}}_{2}+{{{{{\rm{Ca}}}}}}({{{{{\rm{H}}}}}}{}^{12}{{{{{\rm{CO}}}}}}_{3})^{+}$$
(4)

therefore, because of the continuous dissolution/reprecipitation process controlled by Eq. 3, the carbonate is effective in buffering the isotopic signature of CO2 even after a very short time, despite the fact that CO2 is produced mainly by irreversible oxidation of graphite. Equation 1, however, predicts that aragonite buffering is possible only for relatively low CO2/aragonite ratios, ideally below 0.46 where the 13C abundances would be >50%. We will show below that Eq. 1 can be used to evaluate how the isotopic compositions of CO2 evolved from the sedimentary slab at subarc conditions can meet the global average arc signature according to those variables that control the CO2 release, i.e. the fluid/rock ratios and the redox conditions imposed by the environment.

A mathematical model for δ13C of CO2 produced in the sedimentary slab and comparison with global average arcs

Thermodynamic considerations require that at graphite saturation and at fixed PT conditions the amount of CO2 produced by oxidation of graphite in a pure, carbonate-free C–O–H system depends solely on: (i) redox conditions—the more oxidizing they are, the higher the XCO2 (=CO2,aq/(H2O+CO2,aq))34, and (ii) the amount (number of moles) of H2O interacting with graphite—the higher it is, the higher the amount of CO2 at fixed XCO230,35. Therefore, at graphite saturation conditions and as long as graphite is present, the amount of CO2 produced by oxidation is not dependent on the amount of graphite in the system.

In graphite + aragonite + H2O systems, we demonstrated that the carbon contribution to fluids due to aragonite dissolution is negligible at subarc PT conditions compared to graphite, so the amount of CO2 evolved still depends on (i) and (ii) above, with the additional consequence that the graphite/carbonate ratio is again not relevant. Moreover, the carbon content in these fluids is basically ascribable to their CO2 content, as Ca(HCO3)+ is limited to about 1%. However, we noticed a difference in the aragonite + graphite system compared to the graphite-alone (COH) system: the halving of measured XCO2 (=0.166) compared to graphite-saturated COH fluid models (=0.303; Fig. 2a). By analogy with other systems, we suggest that this is likely related to a modification of water activity15, in this case due to aragonite dissolution. Applying the experimentally derived correction factor of 0.548 (=0.166/0.303), we calculated for the aragonite + graphite system the XCO2 values predicted by conventional graphite-saturated fluid thermodynamic model29,30 at different redox conditions (ΔFMQ). The calculated XCO2 were used to calculate the absolute amount of CO2 produced by oxidation of graphite, obtained by fixing the absolute amount (moles) of H2O in the system. Assuming 1 mol aragonite in the system, fixed H2O values correspond also to water/aragonite molar ratios (WAR) and CO2 values to CO2/aragonite molar ratios. Using these last values and Eq. 1, we calculated the 13C abundance of CO2 produced by oxidation of 12C-graphite after interaction with 13C-aragonite. Because in our experimental model 12C-graphite and 13CaCO3 represent analogues of, respectively, ocean organic matter with δ13C ≈ −20 and marine carbonates with δ13C ≈ 0, we are now able to translate 13C abundances (ranging from 0 to 100%) to equivalent δ13C values (ranging from 0‰ to –20‰) using a conventional linear model:

$${\delta }^{13}{{{{{\rm{C}}}}}}\textperthousand =0.2{\times }^{13}{{{{{\rm{C}}}}}} \% {-}20$$
(5)

the calculated δ13C‰ of CO2 repeated for 10,000 different random combinations of the parameters ΔFMQ and WAR have been fitted with a three-variate parametric equation to provide, with an estimated average uncertainty of ±0.2‰, the following mathematical model (Fig. 4):

$${{{{{\rm{\delta }}}}}}{}^{13}{{{{{\rm{C}}}}}}_{{{{{{{\rm{CO}}}}}}}_{2}}{{\textperthousand }}=-20+20\frac{1}{1+{e}^{-f(\triangle {{{{{\rm{FMQ}}}}}},{{{{{\rm{WAR}}}}}})}}$$
(6)

where the function f(ΔFMQ, WAR) is a third-order polynomial found to minimize residuals (parameters and associated uncertainties available in Supplementary Table 7). The global arc δ13C ranging from –2.8 to –3.3 ‰4 (orange box in Fig. 4) intersects the surface of Eq. 6 providing a continuous layer that can be averaged and expressed with the following equation:

$$\Delta {{{{{\rm{FMQ}}}}}}=\frac{2}{1+{e}^{(A+B\times {{{{{\rm{WAR}}}}}}+C{\times {{{{{\rm{WAR}}}}}}}^{2}+D\times {{{{{{\rm{WAR}}}}}}}^{3})}}$$
(7)

equation 7, where A = –0.314(9), B = 2.75(5), C = –1.46(6) and D = 0.37(2), represents the locus of points of all the possible combinations of ΔFMQ and water/aragonite ratio that result in the isotopic composition of sediment-derived CO2 satisfying the global average arc values.

Fig. 4: Mathematical model showing the predicted δ13C of CO2 as a function of both ∆FMQ and water–aragonite (≈ fluid–rock) molar ratio.
figure 4

Modelled CO2 is originated from oxidative dissolution of graphite and interacts with aragonite and aqueous fluids at 3 GPa and 700 °C. Brown surface color: δ13C buffered by graphite. Blue surface color: δ13C buffered by aragonite. Orange box: global average δ13C of arc CO24; the intersection with the mathematical model shows that arc CO2 signatures are met for a wide range of ΔFMQ and fluid–rock ratios. Preferred model (yellow field) corresponds to the intersection among our mathematical model, the global average arc δ13C and typical fluid–rock ratios at slab-top conditions assuming percolating flux (green field; 0.05–0.33 molar = 0.01–0.06 mass). In this case, ΔFMQ ranging from +0.78 to +1.08 are derived for the source of carbon in the sedimentary slab at subarc conditions.

Comparison with conventional mass-balance/equilibrium fractionation models

The observed carbon isotopic exchange among CO2, graphite and aragonite is potential of interest for subducted sediments buried in a wide range of basins; the only limiting factor is the coexistence of calcium carbonate with elemental carbon (formerly organic matter) in whatever proportion and the presence of an aqueous fluid. Our experiments at 3 GPa and 700 °C show that CO2 produced by oxidative dissolution of 12C-graphite (13C = 1.1%; equivalent δ13C = –19.8, cf. Eq. 5 and Supplementary Table 1) becomes rapidly enriched in 13C ending after 240 h with 13C = 82.2% (equivalent δ13C = –3.56) due to its isotopic exchange with 13C-aragonite (13C = 99.6%; equivalent δ13C = –0.08), despite the overwhelming abundance of graphite in the starting materials (graphite/aragonite(molar) = 5.6; see below).

Equivalent δ13C values can be compared with conventional models combining mass-balance calculations with isotope equilibrium fractionation (see Methods). Let’s assume a mixture similar to what investigated experimentally: aragonite (δ13C = 0‰, by analogy with a marine carbonate source) + graphite (δ13C = –20‰; by analogy with a marine organic matter source), equilibrating at 700 °C (pressure effect ignored). Assuming complete solid–solid isotopic exchange and neglecting devolatilization, the final isotopic composition would be a function of the molar ratio between graphite and aragonite, for instance: 1) δ13Cgraphite = –5.29‰; δ13Caragonite = –0.14‰ for graphite/aragonite = 0.01; 2) δ13Cgraphite = –12.57‰; δ13Caragonite = –7.43‰ for graphite/aragonite = 1.0; 3) δ13Cgraphite = –17.77‰; δ13Caragonite = –12.63‰ for graphite/aragonite = 5.6. With the latter abundance value, comparable with the experimental setup, calculations show that the isotopic composition of all the phases in equilibrium would vary slightly as a function of the amount of CO2 in the system produced by oxidation of graphite (Supplementary Fig. 3a), with δ13CCO2 = –10.17‰ for CO2/aragonite = 0 and δ13CCO2 = –11.33‰ for CO2/aragonite = 1. This conventional model is therefore not adequate to reproduce our experimental system, characterized by CO2/aragonite ≈ 0.1, where both aragonite and CO2 converge to much heavier compositions. However, the fit between the isotopic model and the experimental trend improves significantly (Supplementary Fig. 3a) if we assume (1) graphite as chemically reactive but isotopically inert phase and (2) starting CO2 displaying δ13C = –20‰, i.e. the same value of the starting graphite. These assumptions, deduced on the basis of our experimental findings, underline again the need for experimental constraints on deep isotopic exchange processes involving fluids. For instance, an implication of our study could throw some light on the hard debate on the genesis of diamonds characterized by light, organic matter-like isotopic compositions19. In fact, even if we predict that, after having interacted at subarc conditions with graphite-saturated fluids, sedimentary carbonates will be only slightly depleted in 13C, showing the same δ13C −2.8 to −3.3‰ characterizing CO2, they could become locally lighter (cf. brown in Fig. 4) in case of highly oxidized conditions36 or very high water/carbonate (≈fluid/rock) ratios37. Because of their stability at high-pressure conditions38, these carbonates could be subducted further, eventually making available a 12C-enriched source of carbon, which comes from carbonate and not organic matter, in the diamond stability field39 (Fig. 5).

Fig. 5: Carbon isotope exchange at the subarc slab-mantle interface.
figure 5

Organic (brown) and inorganic (dark blue) carbon occurring in open-ocean sediments enters the subduction channel, where it interacts with aqueous fluids produced by dehydration of the down-going oceanic lithosphere. While calcium carbonate dissolution increases almost linearly with depth14, resulting in fluids containing ~5000 ppm C mainly as Ca(HCO3)+ at subarc depth, the oxidative dissolution of graphite increases exponentially and reaches its maximum at subarc conditions, where fluids coexisting with graphite contain ~140,000 ppm C mainly under the form of CO2(aq). With the simultaneous presence of aragonite and graphite, experiments show that the carbon concentration in aqueous fluids is halved, but still controlled by oxidation of graphite to CO2. Nevertheless, as aragonite is very prone to dissolution-reprecipitation processes, it shows an outstanding capacity to buffer the isotopic signature of graphite-derived CO2 to carbonate-like values characterizing the global arc CO2 (orange field). In order to exert its buffering capacity, graphite-derived CO2 must be low with respect to carbonate, which is the case for low fluid–rock ratios coupled with oxidizing conditions (preferred model) or for high fluid–rock ratios coupled with relatively reducing conditions. We suggest that, as a mobile but reactive fluid-phase, CO2 is particularly effective in metasomatizing the supra-subduction mantle (light blue field), eventually imposing its isotopic signature to the deep source of arc magmatism.

The link between sedimentary slab fluids and arc emissions

As an aqueous fluid-phase species, CO2 originated from subducted sediments close to the slab-mantle interface is particularly effective in metasomatizing the subarc supra-subduction mantle, prompting carbonation reactions20,21,23,40 that may facilitate diapiric upwelling and melting processes20,41. Therefore, there is good reason to believe that sediment-derived CO2 can impose its isotopic signature on the subarc mantle, the source of arc magmatism (Fig. 5). The interaction of subducted ocean carbonates and organic matter with fluids coming from the dehydration of the down-going slab25,42 is a continuous process starting early in the forearc region9,43 (Fig. 5; Supplementary Fig. 1). However, our study indicates that the production of CO2 is strictly connected to the stability of graphite in aqueous fluids. At forearc conditions graphite is poorly soluble by oxidative dissolution, even considering the increased dissolution susceptibility for disordered graphitic carbon, in the absence of intense flushing by aqueous fluids (i.e. channelized flow)16,35. Aqueous fluids reach their maximum CO2 contents at subarc conditions (Fig. 5; Supplementary Fig. 1), which are the conditions investigated in our study16. Beneath arcs, the interaction between aqueous fluids coming from the dehydrating oceanic slab and percolating into the overlying carbonate metasediments will thus result in a large amount of CO2 leached out from the slab, as long as graphite is not completely removed by oxidative dissolution, which depends chiefly on the time-integrated fluid flux for given redox state18. Although channelized fluid flow can potentially dissolve the entire amount of organic carbon at blueschist-to-eclogite conditions, the geological record suggests the complete removal is difficult even for very high fluxes and often limited to the more soluble disordered forms of organic carbon18. Fully crystalline graphite, which is characterized by sluggish rates of isotopic diffusion3, is thus expected to persist at subarc depths. We demonstrated experimentally that the organic carbon content is not a relevant variable in controlling the carbon isotopic signature of CO2 produced by dissolution processes, which only depends on redox conditions and fluid-rock ratios. Therefore, δ13C of gaseous arc emissions is not a simple linear combination of the amount of buried organic and inorganic carbon4. One major implication is that fluctuations in δ13CCO2 do not necessarily reflect variations in the source input, for instance, due to shallow crustal carbonate assimilation44,45 or to the burial of anoxic sediments dominated by organic matter19. It could reflect instead local variations in fluid-rock ratios (i.e. channelized vs. percolating flux) or in the redox state (i.e. oxidizing vs reducing conditions). Equation 7 can be used to constrain the average redox state (expressed as ΔFMQ) in the sedimentary carbon source of the slab that meets the global arc δ13CCO2. ΔFMQ values ranging from +0.78 to +1.08 are predicted assuming a fluid–rock mass ratio (≈WAR) of 0.01–0.06 (=0.05–0.33 molar), which has been suggested for infiltrating, non-channelized fluids produced by dehydration reactions in eclogites46,47. These oxygen fugacity values are in agreement with estimations of island-arc-basalt (IAB) sources48,49 and supra-subduction mantle peridotites equilibrated at subarc depths50. It remains an open question how these oxygen fugacity values correlate with the amount of oxygen36 transferred from the slab to the IAB source, with consequences for the redox budget of subarc mantle51.

Method

Experimental approach and characterization of the solid- and fluid phases

In this study, we employed as starting materials: (i) labelled Ca13CO3 (calcite; forming aragonite at run conditions) (Sigma–Aldrich); (ii) synthetic graphite (Sigma–Aldrich), highly ordered as suggested by Raman spectroscopy16; (iii) oxalic acid di-hydrate (Sigma); (iv) MilliQ water, boiled while flushed with N2 to remove dissolved atmospheric CO2. Experiments were buffered using the double-capsule technique52 (Supplementary Fig. 4) with an inner H2-permeable Au60Pd40 capsule, containing the starting materials with the addition of water (~20 wt%), and an outer Au capsule filled with the buffering assemblage fayalite + magnetite + quartz + H2O (FMQ; forming ferrosilite + magnetite + coesite at run conditions, verified by electron microscopy, electron microprobe analysis and Raman spectroscopy; Supplementary Fig. 5). The buffer in the outer capsule constrains directly the fH2 in the inner capsule. The redox conditions (fO2) in the inner capsules are constrained indirectly by the buffer, whose ∆FMQ is +0.76 log units in the outer capsule, to slightly lower fO2 values (∆FMQ = +0.61 log units; see “Thermodynamic modelling” and Supplementary Table 4).

Experiments were performed at 3 GPa and 700 °C using an end-loaded piston-cylinder apparatus. Temperatures were measured with K-type thermocouples and considered accurate to ±5 °C. Pressure calibration is based on the quartz/coesite transition53 and accurate to ±0.01 GPa. Runs with CaCO3 were first pressurized at 3 GPa for 2 h, to promote carbonate crystal growth by cold sintering54 in order to increase the grain size of the synthetic micrometric CaCO3 power. Samples are then heated to 700 °C with a ramp of 100 °C/min. Run durations were from 2.4 to 240 h. Quench is obtained by cutting off the power supply, resulting in a temperature decline of >40 °C per second. CO2, water and other volatiles in the inner capsules were measured quantitatively by quadrupole mass spectrometry (QMS) using the capsule-piercing technique31. The typical analytical uncertainty is 1 mol% for CO2 and H2O. Solid phases were checked by scanning electron microscopy, electron microprobe analyses and micro-Raman spectroscopy.

Isotopic analysis of CO2

The quantitative analysis and the determination of the 13C/12C ratio of CO2 have been performed simultaneously by means of QMS31, monitoring the m/z channel 45 (13CO2) in addition to channels considered for routine analysis. The calibration curve linking the 13C/12C ratio of CO2 and the ratio of the integrated peaks of channels 45 (13CO2) and 44 (12CO2) (44/45 ratio) has been derived by measuring 3 different mixtures with known 13C/12C ratio prepared to start from (i) regular oxalic acid di-hydrate (OAD; Sigma–Aldrich), used also for OAD + Ca13CO3 experiments and (ii) isotopically nearly pure 13C oxalic acid di-hydrate (Sigma–Aldrich), thermally decomposed to CO2 at 250 °C31 (Supplementary Table 8). Analyses performed by QMS are affected by a very small mass-bias effect. Actually, the regression line provided a very low correction factor of 1.03 for the integrated peak of channel 45.

Isotopic analysis of solids: NanoSIMS

13C/12C ratios of carbonate and graphite grains in the experimental sample were determined by nanoscale secondary ion mass spectrometry (NanoSIMS). Measurements were conducted on the Cameca NanoSIMS 50 installed at Muséum National d’Histoire Naturelle of Paris. The sample capsule was included in pure indium and gold-coated (20 nm thick). Secondary ions of 12C12C and 13C12C were collected in multicollection mode to obtain 13C/12C ratios after calibrating with natural abundance graphite and calcite samples. 16O secondary ions were used to locate graphite versus carbonate grains. Mass resolving power was set at a minimum 10,000, enough to resolve interferences on measured secondary ions. Before each analysis, a 5 × 5 µm2 surface area was initially pre-sputtered for 60 s with a 500 pA primary Cs+ rastering beam, in order to remove the gold coating and reach a sputtering steady-state55. For analyses of graphite grains, the primary beam was set to 1 pA and is was scanned over a surface area of 5 × 5 µm2. Nevertheless, to avoid surface contamination, only ions from the inner 2.6 × 2.6 µm2 regions were collected with the “beam blanking mode”. Each analysis consisted of a stack of 100 cycles, with a duration of 2.048 s each. Similar settings were used to carbonate grains, except that the primary beam was set to 9 pA and each cycle was 8.192 s long. 13C/12C of carbonate grains were also investigated using 12C and 13C secondary ions (mass resolving power set to 9000), showing no significant deviation from the measurements using C2 secondary ions. Due to large isotopic variations observed in these samples, instrumental mass fractionation (of a few per mil) was neglected.

Isotopic analysis of solids: LA-ICP-MS

13C/12C ratios of carbonate and graphite grains in the experimental sample were also determined by LA-ICP-MS. Measurements were carried out over a New Wave UP 266 laser ablation system coupled with a Thermo Fisher ICAP-Q ICP-MS (University of Insubria). In order to find the best compromise between ablated surface and sensitivity, ablation conditions were optimized over a natural abundance calcite sample with a known 12C/13C ratio (see Section “Isotopic analysis of solids: Bulk analyses”). As a result, for each spot determination, 20 shots within 1 s were performed, using a 30 μm diameter circular spot size and adjusting the laser power to obtain a fluence value of around 17 J/cm2. Helium was used as the carrier gas (0.85 L/min). Under these conditions the hole depth (subsequently evaluated by scanning electron microscopy) is, on average, about 20 μm. Analytical accuracies and uncertainties have been evaluated by measuring the following internal standards: (i) natural abundance calcite, characterized isotopically by GasBench isotope ratio mass spectometer (IRMS) in two different laboratories (Florence and Milan), displaying an average 13C abundance of 1.1208(2) %; (ii) synthetic labelled CaCO3 produced by precipitation from a Na213CO3 + Na212CO3 solution treated with CaCl2, with 13C abundance of 43.8% measured by HCl dissolution followed by QMS analysis of evolved CO2 (see “Bulk analyses” below). LA-ICP-MS measurements on these standards have been proven to be accurate to ~2% for natural calcite (standard deviation 5.8%) and ~1% for synthetic 13C–12C calcite (standard deviation 0.7%) (Supplementary Table 9). Analyses of graphite are affected by a higher standard deviation of about 12%, due to decreased ablation efficiency.

Isotopic analysis of solids: Bulk analyses

Bulk analyses of the carbonates contained in the experimental capsules have been carried out by HCl decomposition followed by QMS analysis of evolved CO2 (see “Isotopic analysis of CO2”). To perform this analysis sample capsules coming from LA-ICP-MS determination were placed in a U-shaped glass tube located upstream with respect to the QMS analyser. After 5 outgassing cycles with Ar, 1 mL of HCl (4.5 M) was introduced (from an additional port) in the bottom of the U-tube, where the capsule is located. The tube was then closed (by means of a bypass valve of the gas manifold) for 30 min in order to allow the complete decomposition of all carbonates: after this reaction time the tube was then placed in line with the QMS analyser to determine the composition of evolved CO2. The analysis of the natural calcite standard provided a 13C abundance of 1.13%, which is accurate to 0.8% of the certified value. In addition, this method allowed to retrieve the 13C abundance of the synthesized 13C–12C calcite (43.8%) and of Ca13CO3 used as starting material (99.4%) (Supplementary Table 10).

The isotopic characterization of the graphite used as starting material has been performed using EA-IRMS analysis, which yielded a value of 1.0948(2)% 13C.

Thermodynamic modelling of fluid composition

To retrieve the redox conditions in the double-capsule system, oxygen and hydrogen fugacities have been calculated by conventional thermodynamic modelling (Supplementary Table 4). In the outer capsule, containing the buffering assemblage ferrrosilite + magnetite + coesite + H2O, log (fO2/1 bar)= –13.21 and log (fH2/1 bar) = 2.195 have been calculated at P–T conditions of 3 GPa and 700 °C using the Perple_X package30, considering the thermodynamic dataset of Holland and Powell56 revised by the authors in 2004 (hp04ver.dat) and the equation of state “H–O HSMRK/MRK hybrid” of the routine “fluids”. Then, XCO2 [=CO2/(H2O+CO2)molar] = 0.303 for graphite-saturated fluids has been calculated by fixing log (fH2/1 bar)= 2.195, which is homogeneous in the inner and the outer capsule, using the Perple_X equation of state of Connolly and Cesare29 (cf. refs. 15,16 for other details). This XCO2 corresponds to an inner-capsule log (fO2/1 bar)= –13.36 (ΔFMQ = +0.61 log units), which reflects the redox conditions occurring in the runs bearing graphite. Log (fH2/1 bar)= 2.195 has also been used to calculate the fluid speciation and the pH in our experimental systems, using the Deep Earth Water (DEW) model26,27 (Supplementary Table 3).

Conventional isotopic modelling

The global δ13C of the investigated system can be expressed with the following mass balance:

$${\delta }^{13}{C}_{{{{{{\mathrm{global}}}}}}}={m}_{{{{{{\mathrm{graphite}}}}}}}\times {\delta }^{13}{C}_{{{{{{\mathrm{graphite}}}}}}\_i}+{m}_{{{{{{\mathrm{aragonite}}}}}}}\times {\delta }^{13}{C}_{{{{{{\mathrm{aragonite}}}}}}\_i},$$
(8)

where m represents the molar fractions, and δ13Cgraphite_i and δ13Caragonite_i the initial composition of graphite and aragonite, respectively, –20‰ and 0‰ in our isotopic model.

Assuming that CO2 is produced only through oxidation of graphite (graphite_ox), and that equilibrium conditions occur between CO2 and graphite, the isotopic composition of graphite_ox is:

$${\delta }^{13}{C}_{{{{\mathrm{graphite}}}}\_ox}=(1-f)\times ({\delta }^{13}{C}_{{{{\mathrm{graphite}}}}\_i}{-}{\Delta }_{{{{\mathrm{CO}}}_{2}}{-}{{{\mathrm{graphite}}}}})+f\times {\delta }^{13}{C}_{{{{\mathrm{graphite}}}}\_i}$$
(9)

where \(\Delta_{{\mathrm {CO}}_{2}}\)–graphite is the CO2–graphite equilibrium fractionation factor at 700 °C57 and f is the molar fraction of oxidated graphite, i.e. (mgraphite − mCO2)/mgraphite, ranging from 1 (no CO2 produced) to 0 (graphite fully oxidated to CO2).

The isotopic composition of CO2 produced by graphite oxidation is:

$${\delta }^{13}{C}_{{{{{{\mathrm{CO}}}}}}_{2}}={\delta }^{13}{C}_{{{{{{\mathrm{graphite}}}}}}\_{ox}}+{\Delta }_{{{{{{\mathrm{CO}}}}}}_{2}{-}{{{{{\mathrm{graphite}}}}}}},$$
(10)

so that the lobal δ13C including CO213Cglobal_ox), which in a closed system is numerically identical to δ13Cglobal, can be expressed as:

$${\delta }^{13}{C}_{{{{{{\mathrm{global}}}}}}\_ox} = \, ({m}_{{{{{{\mathrm{graphite}}}}}}}{-}{m}_{{{{{{\mathrm{CO}}}}}}_{2}})\times {\delta }^{13}{C}_{{{{{{\mathrm{graphite}}}}}}\_ox}+{m}_{{{{{{\mathrm{CO2}}}}}}}\times {\delta }^{13}{C}_{{{{{{\mathrm{CO}}}}}}_{2}}\\ + \,{m}_{{{{{{\mathrm{aragonite}}}}}}}\times {\delta }^{13}{C}_{{{{{{\mathrm{aragonite}}}}}}\_i},$$
(11)

Assuming isotopic equilibrium between CO2, graphite_ox and aragonite (Supplementary Fig. 3a), the final compositions are:

$${\delta }^{13}{C}_{{{{\mathrm{aragonite}}}}\_f}\,=\, {\delta }^{13}{C}_{{{{\mathrm{global}}}}\_ox}{-}({m}_{{{{\mathrm{graphite}}}}}\left.{-}{m}_{{{{\mathrm{CO}}}}_{2}}\right)\times {-}{\Delta }_{{{{\mathrm{aragonite}}}}{-}{{{\mathrm{graphite}}}}})\\ {-}({m}_{{{{\mathrm{CO2}}}}}\times {-}{\Delta }_{{{{\mathrm{aragonite}}}}{-}{{{\mathrm{CO}}}_{2}}});$$
(12)
$${\delta }^{13}{C}_{{{{{{\mathrm{graphite}}}}}}\_f}={\delta }^{13}{C}_{{{{{{\mathrm{aragonite}}}}}}\_f}{-}{\Delta }_{{{{{{\mathrm{aragonite}}}}}}{-}{{{{{\mathrm{graphite}}}}}}};$$
(13)
$${\delta }^{13}{C}_{{{{{{\mathrm{CO}}}}}}_{2}\_{f}}={\delta }^{13}{C}_{{{{{{\mathrm{aragonite}}}}}}\_f}{-}{\Delta }_{{{{{{\mathrm{aragonite}}}}}}{-}{{{{{\mathrm{CO}}}}}}_{2}},$$
(14)

In calculations considering graphite as isotopically inert (Supplementary Fig. 3b), we assume that CO2 is produced by the oxidation of graphite, but that CO2 does not equilibrate with graphite getting oxidized.

Therefore, only CO2 and aragonite are freely exchanging isotopes. In this case, mgraphite becomes 0 and δ13CCO2 displays the constant value of –20‰. Therefore, the following equations were used instead:

$${\delta }^{13}{C}_{global}={m}_{CO2}\times 20\textperthousand +{m}_{Aragonite}\times {\delta }^{13}{C}_{{aragonite}{\_}{i}};$$
(15)
$${\delta }^{13}{C}_{{{{{{\mathrm{aragonite}}}}}}{\_}{f}}={\delta }^{13}{C}_{{{{{{\mathrm{global}}}}}}}{-}({m}_{{{{{{\mathrm{CO2}}}}}}}\times {\Delta }_{{{{{{\mathrm{aragonite}}}}}}{-}{{{{{\mathrm{CO}}}}}_{2}}});$$
(16)
$${{\delta }^{13}} {C}_{{{{{{{\mathrm{CO}}}}}}_{2}}\_f}={\delta }^{13}{C}_{{{{{{{{\mathrm{aragonite}}}}}}}}\_f}{-}{\Delta }_{{{{{{{{\mathrm{aragonite}}}}}}}}{-}{{{{{{{{\mathrm{CO}}}}}_{2}}}}}}.$$
(17)