Supplementary Material Final

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Supplementary material to: From peridotite to listvenite –

perspectives on the processes, mechanisms and settings of


ultramafic mineral carbonation to quartz-magnesite rocks

Menzel, M.D.; Sieber, M.J.; Godard, M.

1. Thermodynamic modelling methods

1.1. Fluid infiltration-fractionation models (Fig. 3; Fig. 6; Fig. S1)

Klein and Garrido (2011) used thermodynamic reaction path models to explore listvenite formation
by reaction of harzburgite with CO2-enriched seawater at 100 – 400 °C and 50 MPa pressure,
corresponding to shallow oceanic hydrothermal settings. Similar models for high pressures (0.5 –
1.0 GPa) and metamorphic CO2-bearing fluid have been presented by Kelemen et al. (2022), again
for a harzburgite protolith. Both models are based on a titration approach using the EQ3/6 software,
where either increasing fractions of reacting rock are incrementally titrated into a fluid reservoir or
fluid aliquots are added incrementally to the rock. Here we calculated models of listvenite formation
using the 0-D infiltration-fractionation mode of the Gibbs energy minimization code Perple_X
(Connolly and Galvez, 2018). These provide a different, complementary perspective: rather than
titration, the calculations simulate an idealized chemical flow-through experiment, where in each
calculation increment a CO2-bearing fluid aliquot is added and equilibrated with the rock, after which
the remaining equilibrium fluid (outflow fluid) is removed before addition of another fluid aliquot.
This approach allows for fast exploration of different protolith compositions while accounting for
solid solutions, fluid solvent models and prediction of solute concentrations to high pressures and
temperatures. In comparison to titration-type models, the fluid infiltration-fractionation models are
path-dependent, meaning that e.g. interaction of 100 aliquots of 1 mol fluid does not equal addition
of 100 mol fluid in one aliquot, because loss of solute components in the outflow fluid depends on
the mineral assemblage that changes with reaction progress.

To illustrate carbonation reaction progress for harzburgite (Fig. 3), dunite and lherzolite (Fig. 6) and
komatiite (Fig. S1) we used respective bulk rock compositions that we consider representative for a
wide range of natural ultramafic protoliths, albeit simplified here for the purpose of clarity (Table
S1). We neglect Cr as it is predominantly hosted by Cr-spinel, which is commonly observed as a
relict phase in listvenite and thus mostly does not equilibrate with the bulk rock composition. Ca was
also not considered in dunite and harzburgite calculations as typically small or negligible Ca contents
in these rocks are prone to become fully depleted from the rock composition in the infiltration model
due to fluid fractionation. For simplicity, we chose here an idealized infiltrated fluid composed only
of H2O – CO2 with the arbitrary choice of XCO2 = 0.029, which is high enough for listvenite formation
at the model conditions of 300 °C and 0.5 GPa but not unreasonably CO2-rich.
For the infiltration calculations we use the ds6.2 thermocalc database for minerals (Holland and
Powell, 2011), the HKF formulation (Helgeson et al., 1981) and the deep earth water model (DEW,
2019 version) aqueous database (Sverjensky et al., 2014) (DEW19HP622ver_elements database in
Perple_X). Solid solution models were applied for olivine, magnesite-siderite and dolomite-ankerite
(all from Holland and Powell, 1998), ortho- and clinopyroxenes (Holland and Powell, 1996),
antigorite (Padrón-Navarta et al., 2013), chlorite (Holland et al., 1998), and ideal solution models for
tremolite, talc, magnetite and brucite. For simplicity, hematite, andradite, pyrophyllite, graphite,
calcite and quartz were considered as pure phases. Note that we used a set of comparatively simple
and well-established solution models that we consider sufficient to serve the main purpose of the
calculations, but that more detailed investigations should take into account newer and more complex
solution models, in particular for spinel, amphibole, chlorite and antigorite (e.g., Eberhard et al.,
2023). The solvent and solute composition of the equilibrated outflow fluid is calculated using the
COH-Fluid model from Connolly and Galvez (2018), with the equations of state for H2O and CO2
after Pitzer and Sterner (1995) and after Jacobs and Kerrick (1981) for CH4. The use of a solvent
COH-Fluid model for neutral molecular fluid species H2O, CO2, CO, CH4 and H2, combined with
lagged speciation of aqueous solutes (Connolly and Galvez, 2018) allows to account for CO2
concentrations that are beyond the limits of the HKF approach. Accordingly, we excluded in the
calculation neutral/molecular fluid species of the DEW model, such as CO2,aq, H2CO3,aq, CO,aq, H2,aq
and methane. We further did not consider more complex aqueous species that have been primarily
derived to fit solute concentrations in high-pressure (3 – 5 GPa) peridotite experiments (MgSiC+ and
Fe(HCOO)) (Dvir et al., 2013; Tiraboschi et al., 2017). The role of MgO,aq, which we did include in
the calculations, is unclear. This is due to an unresolved discrepancy of more than one order of
magnitude of Mg solute concentration between classical weight-loss dissolution experiments (Macris
et al., 2020; Newton and Manning, 2002) and cryo diamond trap experiments (Tiraboschi et al., 2017)
at the same conditions. Excluding MgO,aq from the infiltration models leads to minor differences in
the chemistry of the outflow fluid and consequently in the compositional evolution at very high fluid
flux (high integrated fluid/rock ratios), but does not significantly affect the interpretation and
conclusions obtained from our modelling results here.

1.2. Classical T–XCO2 modelling at variable P (Fig. 7)


To illustrate the influence of pressure and temperature on the minimum XCO2 required for different
carbonated assemblages (Fig. 7), we calculated classical T-XCO2 pseudosections at variable pressures
in the MgO-FeO-SiO2 system with saturated H2O-CO2 fluid, using the standard gridded minimization
mode in Perple_X (Connolly, 2005, 2009). In this calculation we used a simplified harzburgite
composition (in wt%: MgO: 39.21, FeO: 4.19, SiO2: 38.77) and applied a logarithmic scaling for
XCO2 for better visualization. For consistency with the infiltration models, we applied the hp622ver
database (Holland and Powell, 2011) with the same solid solution models for olivine, orthopyroxene,
magnesite, antigorite, talc and brucite (c.f. Appendix section 1.1), and equations of state for H2O and
CO2 after Pitzer and Sterner (1995). Discrepancies with calculations using the CORK equation of
state (Holland and Powell, 1991) and the internally consistent database by Holland and Powell (1998)
were negligible at the model conditions.

1.3. CO2 fluid concentration of source rock vs. listvenite in dependence of P–T (Fig. 12)

To assess the potential of CO2-bearing fluid released from meta-sediment with respect to ultramafic
rock carbonation at variable pressure and temperature, we calculated P–T pseudosections (2D gridded
minimization in Perple_X) of graphite-calcschist, soapstone and listvenite, with a particular focus on
the composition of COH-fluid in equilibrium with the respective metamorphic assemblages. We
chose here the composition of a garnet-bearing calcite-dolomite quartz mica schist as an example
(Table S2), to which we arbitrarily added a small amount of elemental carbon since fluids equilibrated
with impure graphite-carbonate-silicate rocks can have particularly high CO2 concentrations. Similar
impure carbonate-silicate rocks are likely equally suitable CO2 sources. From the pseudosections we
then calculated the difference of the CO2 concentration between fluid of graphite-calcschist and of
that in equilibrium with antigorite-talc-magnesite (soapstone) and quartz-talc-magnesite (listvenite)
assemblages (Fig. 12). This provides a first-order indication of the extent of carbonation that may be
achieved from adjacent meta-sediment devolatilization at different P–T conditions. Since we applied
the DEW data for electrolytic COH-Fluid as in the infiltration calculations, the models also predict
the concentrations of carbon-bearing solute species, for instance NaHCO3,aq. However, the effect of
electrolyte species is not straightforward, because species like NaHCO3,aq will also be present in the
outflow fluid after carbonation. Due to the absence of Na-bearing solids in soapstone/listvenite, this
effect is not feasible to model using P–T pseudosections that we applied here as a first-order
assessment. Therefore, we limited the discussion here to CO2,aq (expressed here as XCO2 for better
comparison with classical T-XCO2 models), obtaining a conservative estimate of the carbonation
potential.
The graphite-calcschist pseudosection was calculated using the DEW19HP62ver_elements database
of Perple_X with solution models appropriate for meta-sediments, including garnet, biotite,
orthopyroxene, mica, cordierite, staurolite, chlorite and chloritoid (all from White et al. (2014)),
calcite, stilpnomelane, pumpellyite and carpholite (Massonne and Willner, 2008), epidote (Holland
and Powell, 2011), feldspar (Fuhrman and Lindsley, 1988), dolomite (Holland and Powell, 1998),
amphibole (Green et al., 2016), omphacite (Green et al., 2007) and COH-Fluid (Connolly and Galvez,
2018). Equations of state of molecular fluid species and excluded aqueous solute species are the same
as in the infiltration models. The use of an omphacite solution model to account for compositional
variability in clinopyroxene suppressed the stability of amphibole at most conditions of the
calculation. Calculations without omphacite solid solution result in more amphibole but also pure
jadeite. The latter increases the NaHCO3,aq concentration in fluid and thereby decreases XCO2 at high-
P and elevated T in comparison to the XCO2 as used in Figure 12. Although an omphacite solution
model may not be fully appropriate for the complete range of P–T conditions of our calculation, we
consider it justified here since omphacitic and Na-Fe3+ bearing clinopyroxene is a common
observation in subducted calcschists (Cook-Kollars et al., 2014; Piccoli et al., 2016).

Soapstone and listvenite compositions were taken from appropriate steps of the harzburgite
infiltration model (Fig. 3), with 1 mol H2O added to attain fluid saturation across the P–T range of
interest (Table S2). Thermodynamic databases, solid solution models and fluid equations of state for
the calculation of the P-T pseudosections of soapstone and listvenite were the same as for the
infiltration models (Appendix section 1.1).

1.4. Listvenite stability during metamorphism (Fig. 13)

To illustrate the predicted stability of listvenite and soapstone upon prograde metamorphism we re-
calculated the P-T pseudosections that form the basis of Figure 12 for a wider range of pressures and
temperatures (Fig. 13). All calculation conditions and chemical system are the same as of
soapstone/listvenite modelling for Fig. 12 (Appendix section 1.3), with the model rock compositions
shown in Table S2. Importantly, the model composition includes an arbitrary amount of excess H2O
so that the molecular and electrolytic composition of fluid can be investigated at all P–T conditions.
This results in lower XCO2 of fluid in fully decarbonated assemblages (no magnesite stable, XCO2
approaching 1) in comparison to models without excess water.
2. Supplementary Tables
Table S1. Model bulk compositions used in infiltration calculations. All rock compositions were
normalized to mol/kg rock on a volatile-free basis and with all Fe as FeO.
mol/kg dunite1 harzburgite2 lherzolite3 komatiite4 fluid
Si 6.8249 7.5593 7.6227 7.6871
Al 0.5686 1.2285
Fe 1.2095 1.1400 1.1663 1.4524
Mg 12.3130 11.3971 9.8437 7.8737
Ca 0.5791 0.9596
C 0.03
H2 1.00
O2 13.5861 13.8279 13.8437 13.7513 0.53
1
simplified dunite based on OM94-52d of Hanghøj et al. (2010)
2
simplified harzburgite based on OM94-52h of Hanghøj et al. (2010)
3
simplified lherzolite based on the average depleted MOR mantle (DMM) after Workman and Hart (2005), with adjusted
Al (assuming Cr-spinel remains a relict phase during carbonation)
4
simplified komatiite based on the average of ultramafic Abitibi greenstone belt komatiites after Barnes (1985)

Table S2. Model bulk compositions used to assess the soapstone and listvenite formation potential
by CO2-bearing fluid release from meta-sediment (Fig. 12) and metamorphic devolatilization of
soapstone and listvenite (Fig. 13).
mol/kg Gph-calcschist1 soapstone2 listvenite3 listvenite (lherz.)4
Si 6.91565 7.53229 7.47827 7.48197
Al 0.82540 0.568573
Fe 0.26848 1.12602 1.11046 1.12597
Mg 0.16369 11.0958 11.0670 9.77862
Ca 4.79432 0.557604
Na 0.20652
K 0.10610
C 5.06342 4.29689 8.59277 10.2349
H2 0.92099 4.66848 2.03145 1.71142
O2 15.6897 20.3974 23.2981 24.8568
1
composition of garnet-bearing calcschist SN-30 in Menzel et al. (2019), with arbitrary 0.1 mol graphite added. The O2
amount accounts for measured Fe3+/Fe2+ in the sample.
2
soapstone composition obtained from the harzburgite infiltration model after 150 mol fluid infilftration (Fig. 3), with 1
mol H2O added to attain fluid saturation across the P–T range of interest.
3
listvenite composition obtained from the harzburgite infiltration model after 300 mol fluid infilftration (Fig. 3), with 1
mol H2O added to attain fluid saturation across the P–T range of interest.
4
listvenite composition obtained from the lherzolite infiltration model after 460 mol fluid infilftration (Fig. 6b), with 1
mol H2O added to attain fluid saturation across the P–T range of interest. Only used here for Figure S2.
3. Supplementary Figures

Figure S1. From top to bottom: Evolution of phase proportions, outflow fluid composition (selected
species), ΔpH and cumulative mass and volume changes with increasing fluid-rock interaction
during carbonation of komatiite (composition see Table S1) from infiltration modelling (P = 0.5
GPa, T = 300 °C, infiltrating fluid XCO2 = 0.029). Abbreviations see Fig. 6b.
Figure S2. Molar Mg/Si and Ca/Mg in fluid in equilibrium with listvenite (lherzolitic protolith
composition, see Table S2) as a function of pressure and temperature. White dashed lines mark
elemental ratios of 1; the yellow dotted lines highlight the switch from alkaline to acidic pH
(relative to neutral pH at the respective P–T). In the white area at high-T/low-P, fluid is extremely
CO2-rich such that the lagged speciation calculation failed to reliably predict the solute chemistry.
Figure S3. P–T pseudosection of graphite-bearing garnet-calcschist (model bulk composition see
Table S2), from which the CO2 concentration of in COH-fluid was calculated to compare source
fluid XCO2 with soapstone and listvenite formation potential (Fig. 12). For illustration, some small
low-variance fields are omitted or not labelled here. Mineral abbreviations: Anl: andalusite; Stlp:
stilpnomelane; Wm: white mica (several co-existing micas of different composition occur at some
conditions); Dol: dolomite; Cal: calcite; Arg: aragonite; Mgs: magnesite; Fsp: feldspar (several co-
existing feldspars of different composition occur at some conditions); Chl: chlorite; Ep: epidote; Bt:
biotite; Cpx: clinopyroxene (mostly acmitic to jadeitic here); Am: amphibole; Lws: lawsonite.
Quartz, graphite (at lowT/highP: diamond) and COH-Fluid occur in all fields.
Figure S4. CO2 concentration in COH-Fluid (expressed as XCO2) in dependence of P–T in
soapstone, listvenite and graphite-calcschist.

4. References
Barnes, S.-J., 1985. The petrography and geochemistry of komatiite flows from the Abitibi
Greenstone Belt and a model for their formation. Lithos 18, 241-270.

Connolly, J.A.D., 2005. Computation of phase equilibria by linear programming: A tool for
geodynamic modeling and its application to subduction zone decarbonation. Earth and Planetary
Science Letters 236, 524-541.

Connolly, J.A.D., 2009. The geodynamic equation of state: What and how. Geochemistry,
geophysics, geosystems 10, Q10014.

Connolly, J.A.D., Galvez, M.E., 2018. Electrolytic fluid speciation by Gibbs energy minimization
and implications for subduction zone mass transfer. Earth and Planetary Science Letters 501, 90-
102.

Cook-Kollars, J., Bebout, G.E., Collins, N.C., Angiboust, S., Agard, P., 2014. Subduction zone
metamorphic pathway for deep carbon cycling: I. Evidence from HP/UHP metasedimentary rocks,
Italian Alps. Chemical Geology 386, 31-48.

Dvir, O., Angert, A., Kessel, R., 2013. Determining the composition of C-H-O liquids following
high-pressure and high-temperature diamond-trap experiments. Contributions to Mineralogy and
Petrology 165, 593-599.

Eberhard, L., Frost, D.J., McCammon, C.A., Dolejš, D., Connolly, J.A.D., 2023. Experimental
Constraints on the Ferric Fe Content and Oxygen Fugacity in Subducted Serpentinites. Journal of
Petrology 64, egad069.

Fuhrman, M.L., Lindsley, D.H., 1988. Ternary-feldspar modeling and thermometry. American
Mineralogist 73, 201-215.
Green, E., Holland, T., Powell, R., 2007. An order-disorder model for omphacitic pyroxenes in the
system jadeite-diopside-hedenbergite-acmite, with applications to eclogitic rocks. American
Mineralogist 92, 1181-1189.

Green, E.C.R., White, R.W., Diener, J.F.A., Powell, R., Holland, T.J.B., Palin, R.M., 2016.
Activity–composition relations for the calculation of partial melting equilibria in metabasic rocks.
Journal of Metamorphic Geology 34, 845-869.

Hanghøj, K., Kelemen, P.B., Hassler, D., Godard, M., 2010. Composition and Genesis of Depleted
Mantle Peridotites from the Wadi Tayin Massif, Oman Ophiolite; Major and Trace Element
Geochemistry, and Os Isotope and PGE Systematics. Journal of Petrology 51, 201-227.

Helgeson, H.C., Kirkham, D.H., Flowers, G.C., 1981. Theoretical prediction of the thermodynamic
behavior of aqueous electrolytes by high pressures and temperatures; IV, Calculation of activity
coefficients, osmotic coefficients, and apparent molal and standard and relative partial molal
properties to 600 degrees C and 5kb. American Journal of Science 281, 1249-1516.

Holland, T., Baker, J., Powell, R., 1998. Mixing properties and activity-composition relationships
of chlorites in the system MgO-FeO-Al2O3-SiO2-H2O. European Journal of Mineralogy 10, 395-
406.

Holland, T., Powell, R., 1991. A Compensated-Redlich-Kwong (CORK) equation for volumes and
fugacities of CO2 and H2O in the range 1 bar to 50 kbar and 100-1600°C. Contributions to
Mineralogy and Petrology 109, 265-273.

Holland, T., Powell, R., 1996. Thermodynamics of order-disorder in minerals; II, Symmetric
formalism applied to solid solutions. American Mineralogist 81, 1425-1437.

Holland, T., Powell, R., 1998. An internally consistent thermodynamic data set for phases of
petrological interest. Journal of Metamorphic Geology 16, 309-343.

Holland, T., Powell, R., 2011. An improved and extended internally consistent thermodynamic
dataset for phases of petrological interest, involving a new equation of state for solids. Journal of
Metamorphic Geology 29, 333-383.

Jacobs, G.K., Kerrick, D.M., 1981. Methane: An equation of state with application to the ternary
system H2O-CO2-CH4. Geochimica et Cosmochimica Acta 45, 607-614.

Kelemen, P.B., Carlos de Obeso, J., Leong, J.A., Godard, M., Okazaki, K., Kotowski, A.J.,
Manning, C.E., Ellison, E.T., Menzel, M.D., Urai, J.L., Hirth, G., Rioux, M., Stockli, D.F., Lafay,
R., Beinlich, A.M., Coggon, J.A., Warsi, N.H., Matter, J.M., Teagle, D.A.H., Harris, M.,
Michibayashi, K., Takazawa, E., Al Sulaimani, Z., the Oman Drilling Project Science Team, 2022.
Listvenite Formation During Mass Transfer into the Leading Edge of the Mantle Wedge: Initial
Results from Oman Drilling Project Hole BT1B. Journal of Geophysical Research: Solid Earth 127,
e2021JB022352.

Klein, F., Garrido, C.J., 2011. Thermodynamic constraints on mineral carbonation of serpentinized
peridotite. Lithos 126, 147-160.

Macris, C.A., Newton, R.C., Wykes, J., Pan, R., Manning, C.E., 2020. Diopside, enstatite and
forsterite solubilities in H2O and H2O-NaCl solutions at lower crustal and upper mantle conditions.
Geochimica et Cosmochimica Acta 279, 119-142.
Massonne, H.J., Willner, A.P., 2008. Phase relations and dehydration behaviour of psammopelite
and mid-ocean ridge basalt at very-low-grade to low-grade metamorphic conditions. European
Journal of Mineralogy 20, 867-879.

Menzel, M.D., Garrido, C.J., López Sánchez‐Vizcaíno, V., Hidas, K., Marchesi, C., 2019.
Subduction metamorphism of serpentinite‐hosted carbonates beyond antigorite‐serpentinite
dehydration (Nevado‐Filábride Complex, Spain). Journal of Metamorphic Geology 37, 681– 715.

Newton, R.C., Manning, C.E., 2002. Solubility of enstatite + forsterite in H2O at deep crust/upper
mantle conditions: 4 to 15 kbar and 700 to 900°c. Geochimica et Cosmochimica Acta 66, 4165-
4176.

Padrón-Navarta, J.A., López Sánchez-Vizcaíno, V., Hermann, J., Connolly, J.A.D., Garrido, C.J.,
Gómez-Pugnaire, M.T., Marchesi, C., 2013. Tschermak's substitution in antigorite and
consequences for phase relations and water liberation in high-grade serpentinites. Lithos 178, 186-
196.

Piccoli, F., Vitale Brovarone, A., Beyssac, O., Martinez, I., Ague, J.J., Chaduteau, C., 2016.
Carbonation by fluid-rock interactions at high-pressure conditions: Implications for carbon cycling
in subduction zones. Earth and Planetary Science Letters 445, 146-159.

Pitzer, K.S., Sterner, S.M., 1995. Equations of state valid continuously from zero to extreme
pressures with H2O and CO2 as examples. International Journal of Thermophysics 16, 511-518.

Sverjensky, D.A., Harrison, B., Azzolini, D., 2014. Water in the deep Earth: The dielectric constant
and the solubilities of quartz and corundum to 60 kb and 1200 °C. Geochimica et Cosmochimica
Acta 129, 125-145.

Tiraboschi, C., Tumiati, S., Sverjensky, D., Pettke, T., Ulmer, P., Poli, S., 2017. Experimental
determination of magnesia and silica solubilities in graphite-saturated and redox-buffered high-
pressure COH fluids in equilibrium with forsterite + enstatite and magnesite + enstatite.
Contributions to Mineralogy and Petrology 173, 2.

White, R.W., Powell, R., Holland, T.J.B., Johnson, T.E., Green, E.C.R., 2014. New mineral
activity–composition relations for thermodynamic calculations in metapelitic systems. Journal of
Metamorphic Geology 32, 261-286.

Workman, R.K., Hart, S.R., 2005. Major and trace element composition of the depleted MORB
mantle (DMM). Earth and Planetary Science Letters 231, 53-72.

You might also like