You are on page 1of 15

Available online at www.sciencedirect.

com

ScienceDirect
Geochimica et Cosmochimica Acta 272 (2020) 21–35
www.elsevier.com/locate/gca

Magnesium isotope fractionation during hydrothermal


seawater-basalt interaction
Martin Voigt a,b,⇑, Christopher R. Pearce c, David M. Fries c, Andre Baldermann d
Eric H. Oelkers a,e
a
Géosciences Environnement Toulouse (GET) – CNRS, 14 Avenue Édouard Belin, 31400 Toulouse, France
b
Institute of Earth Sciences, University of Iceland, Sturlugötu 7, 101 Reykjavı́k, Iceland
c
National Oceanography Centre Southampton, University of Southampton Waterfront Campus, European Way, Southampton SO14
3ZH, United Kingdom
d
Institute of Applied Geosciences & NAWI Graz Geocenter, Graz University of Technology, Rechbauerstr. 12, 8010 Graz, Austria
e
Earth Sciences, UCL, Gower Street, London WC1E 6BT, United Kingdom

Received 19 April 2019; accepted in revised form 21 December 2019; available online 31 December 2019

Abstract

Fluid-rock interactions in hydrothermal systems at or near mid-oceanic ridges (MOR) play a major role in determining the
composition of the oceanic crust and seawater. To quantify the processes that govern cation exchange in these environments
we have experimentally studied the isotopic evolution of d26/24Mg in the fluid phase during seawater-basalt interaction at 250
and 290 °C. Mass balance constraints indicate that isotopically heavy Mg was preferentially incorporated into non-
exchangeable (octahedral) sites in secondary clay minerals such as saponite (Mg-rich smectite), leaving residual fluids enriched
26=24
in light Mg isotopes. The magnitude of fractionation observed during smectite precipitation in our experiments (eSmectiteLiquid )
ranged from 0:35‰ to 0:42‰. This observation, which contrasts with the preferential uptake of light Mg isotopes into bio-
genic and inorganic marine carbonates, highlights the potential utility of Mg isotopes as tracers of the precipitation dynamics
of authigenic Mg-silicate and Mg-carbonate phases. Furthermore, although Mg isotopic fractionation is often masked by the
almost complete removal of Mg in high temperature marine hydrothermal systems, our experiments demonstrate that it does
become significant at lower temperatures where Mg removal by clay formation is incomplete. Under such conditions, this
fractionation will create isotopically light fluids due to smectite precipitation, thus potentially represents an important com-
ponent of the marine Mg isotope inventory.
Ó 2019 Elsevier Ltd. All rights reserved.

Keywords: Mid-ocean ridges; Ocean chemistry; Geochemical cycles; Stable isotopes

1. INTRODUCTION that fractionate its three stable isotopes (24Mg, 25Mg and
26
Mg). Rivers, which transport Mg derived from the weath-
The Mg isotopic composition of seawater is controlled ering of continental carbonate and silicate rocks, are
by a combination of inputs and outputs from reservoirs thought to be the major source of Mg to the oceans, while
with distinct isotopic compositions, and exchange processes seafloor hydrothermal systems and carbonate mineral pre-
cipitation are the dominant sinks (Tipper et al., 2006b;
Higgins and Schrag, 2015; Huang et al., 2018). The study
⇑ Corresponding author at: Institute of Earth Sciences, Univer- of marine evaporite minerals and carbonate-hosting sedi-
sity of Iceland, Sturlugötu 7, 101 Reykjavı́k, Iceland. mentary successions suggest that the Mg concentration of
E-mail address: martinvoigt@hi.is (M. Voigt). seawater varied considerably over the Phanerozoic and is

https://doi.org/10.1016/j.gca.2019.12.026
0016-7037/Ó 2019 Elsevier Ltd. All rights reserved.
22 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

currently increasing (Horita et al., 2002; Fantle and required for mineral formation, and the presence or absence
DePaolo, 2006; Higgins and Schrag, 2012). The current of seed crystals and precipitation inhibitors). Furthermore,
and past state of the marine Mg isotopic system neverthe- fractionation also depends on the physicochemical and sur-
less remains unclear, as Mg isotope records derived from face characteristics of the alteration mineral (such as crystal
foraminifera from the past 40 Ma indicate a rapid decrease structure and bond lengths, specific surface area, surface
in seawater d26/24Mg over the past 15 Ma (Pogge von charge, etc., Li et al., 2014; Baldermann et al., 2018). Such
Strandmann et al., 2014), whereas pelagic marine carbon- factors can be further complicated by aqueous solution spe-
ates from the past 80 Ma suggest only small changes in sea- ciation, as Mg isotope fractionation can occur between
water d26/24Mg over this period (Higgins and Schrag, 2015). individual species (e.g., Schott et al., 2016).
The interpretation of such changes (or lack thereof) is chal- To better constrain the behaviour of Mg isotopes in
lenging as the influence of the different fluxes on the Mg iso- marine hydrothermal systems, we performed a series of
tope budget remains uncertain. In particular the extent of experiments reacting glassy and crystalline basalt with sea-
isotopic fractionation occurring in marine hydrothermal water at temperatures of 250 °C and 290 °C. These temper-
systems is poorly constrained at present (Pogge von atures are intermediate between the low-temperature
Strandmann et al., 2014; Higgins and Schrag, 2015). marine hydrothermal systems at <100 °C and the high-
Although basaltic rock dissolution in hydrothermal systems temperature axial hydrothermal systems that produce fluids
temporarily releases Mg to the fluid phase, this release is of >400 °C and, where necessary to (1) achieve sufficiently
often counterbalanced by Mg-rich clay mineral precipita- fast reaction progress to study this isotopic exchange in
tion. The net result of this coupled process is the almost the laboratory, and (2) to avoid extremely rapid removal
complete removal of Mg from the seawater in high- of Mg from the seawater into secondary phases at even
temperature hydrothermal systems with temperatures higher temperatures. The two rock types were chosen to
exceeding 400 °C, as well as significant Mg removal at explore the potential impact that the presence of distinct
lower temperature conditions (Mottl and Wheat, 1994; mineral phases had on the chemical and isotopic evolution
Elderfield and Schultz, 1996). Consequently, there is a net of the marine basaltic crust relative to the more reactive
transfer of Mg from seawater to the oceanic crust in such glassy components. Unlike previous experimental studies
settings (e.g., Elderfield and Schultz, 1996; Coogan and of such systems (e.g., Bischoff and Dickson, 1975;
Dosso, 2012; German and Seyfried, 2014), and in the case Seyfried and Bischoff, 1977; Mottl and Holland, 1978;
of non-complete Mg removal (especially at lower tempera- Seyfried and Bischoff, 1979; Mottl et al., 1979; Seyfried
tures) this process potentially affects the oceanic Mg isotope and Bischoff, 1981; Seyfried and Mottl, 1982; Seewald
budget (Pogge von Strandmann et al., 2014; Higgins and and Seyfried, 1990), the temporal evolution of the aqueous
Schrag, 2015). Mg isotopic composition was followed during the experi-
Although the number of studies of Mg isotopes in ments via regular sampling, enabling the behaviour of Mg
altered oceanic basalts to date are limited (Teng, 2017; isotopes during the dissolution of basalt and the formation
Huang et al., 2018; Guo et al., 2019), current data suggests of secondary alteration minerals to be constrained.
a wide range of Mg isotopic compositions can result from
carbonate and clay mineral precipitation during low- 2. METHODS
temperature seawater alteration. A relatively strong prefer-
ential uptake of light Mg isotopes is commonly observed by 2.1. Experimental design
Mg-containing carbonate mineral precipitation in shelf
regions, such as during the dolomitization of ancient plat- The experiments conducted in this study were designed
form carbonates (Galy et al., 2002; Tipper et al., 2006b; to investigate chemical exchange during seawater-basalt
Higgins and Schrag, 2010; Teng et al., 2010; Pearce et al., interaction by reacting two types of basalt with different
2012; Mavromatis et al., 2013); this process has been sug- crystallinity with seawater in rocked closed-system Ti-
gested to be the main driver of past changes in seawater reactors at 250 and 290 °C for 53 and 92 days, respectively
d26/24Mg (Pogge von Strandmann et al., 2014). On the other (Fig. 1, Table 1). The basalt used consisted of either basaltic
hand, Mg-bearing silicate alteration minerals such as clays glass or crystalline powders derived from Stapafell in SW
show a smaller degree of Mg isotopic fractionation, with Iceland (40–80 lm size fraction, obtained by milling, siev-
most studies suggesting the preferential incorporation of ing and subsequent cleaning from ultra-fine grains by
heavy Mg isotopes into the non-exchangeable sites (i.e. repeated ultrasonic cleaning in ultrapure water). The speci-
octahedral sheets) of secondary clay minerals (Tipper fic surface areas of the glass and crystalline powders were
et al., 2006a; Higgins and Schrag, 2010; Teng et al., 2010; 0.1 and 0.4 m2/g, respectively, as determined by the BET
Wimpenny et al., 2014; Ryu et al., 2016). There is also technique (Brunauer et al., 1938) on a Quantachrome Auto-
evidence for Mg isotope fractionation in the opposite sorb 1-MP analyser using Kr gas. The chemical composi-
sense - the preferential incorporation of lighter Mg isotopes tion of the samples, previously reported by Oelkers and
into the exchangeable interlayer sites of Mg-bearing clay Gı́slason (2001) and Gudbrandsson et al. (2011), is pro-
minerals (Pogge von Strandmann et al., 2008; Wimpenny vided in Table 2 and is nearly identical to that of mid-
et al., 2010). This suggests that the metal ion (Mg) uptake ocean ridge basalt (MORB, which, on average, contains
by secondary clay minerals and the related isotope fraction- slightly less Mg and Fe but more Si and Al compared to
ation mechanism are highly dependent on the starting fluid the basalt used in this study; White and Klein, 2014). Both
composition (including the pH, concentration of elements powders were reacted with natural seawater collected from
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 23

end of the experiments. The pressure in the reactors, which


Rocking was constrained by equilibrium between liquid and gas, was
motion
Furnace approximately 39 bar and 73 bar at 250 °C and 290 °C,
respectively, as calculated using thermodynamic data for
the H2O-NaCl system of Driesner and Heinrich (2007)
e
lav

Vapor and Driesner (2007). Further details on the fluid sampling


tocu

PTFE gasket
Valve and experimental design used in this study are provided
Ti-a

Seawater Cooling water


Frit
in Voigt et al. (2018b).
Basalt
powder
2.2. Analytical methods
Filter
2.2.1. Aqueous concentrations
Sample The pH of all of the sampled fluids was measured at 22
10 cm
Thermocouple ± 1 °C using a Metrohm microelectrode, calibrated against
Fig. 1. Schematic design of Ti-autoclaves used for the experiments pH buffers of pH 4.0, 6.9, and 9.2. The uncertainty on these
in this study. The Ti-autoclaves were rocked to ensure mixing of pH measurements is ±0.07 pH units, based on replicate
the solids with the seawater. The cooling apparatus attached to the analyses. For the 250 °C experiments, aqueous Na, Mg,
valve allowed fluid samples to be taken during the experimental Si, K, Ca, Sr, Li and Fe concentrations were determined
run. by inductively coupled plasma atomic emission spectrome-
ter (ICP-AES, iCAP6000 Series, ThermoFisher), and by
the equatorial Pacific at a depth of 350 m (EqPac cruise inductively coupled plasma mass spectrometry (ICP-MS,
November 1992; Murray et al., 1997). The seawater was Agilent 7500ce) for the 290 °C experiments. The analytical
acidified upon collection using distilled concentrated uncertainties and detection limits of these analyses are
HNO3 to prevent the precipitation of solids and growth reported in Electronic Supplementary Table 1. For the
of organic material. The pH of the seawater required for 290 °C experiments, Na concentrations were also deter-
the experiments was re-adjusted to its natural value of mined by flame atomic absorption spectrometry (AAS,
8.14 by the gradual addition of a 25% ammonia solution AAnalyst 600, Perkin Elmer). The concentrations deter-
(EMD Millipore Suprapur) prior to use in the experiments. mined for the experimental samples using the different tech-
The mass ratio of seawater to basaltic powder was 10 in niques are consistent within the analytical uncertainties.
all experiments, in line with previous experiments and esti- Additionally, aqueous Al concentrations were determined

mates for submarine hydrothermal systems (Seyfried and by furnace AAS, while aqueous SO2 4 and Cl concentra-
Bischoff, 1979; Seyfried and Bischoff, 1981). Fluid samples tions were measured using ion chromatography (IC, Dio-
(22 mL) were regularly retrieved from the reactor during nex ICS 2000). Additional details of these analytical
the experiments using a sampling valve with an attached methods can be found in Voigt et al. (2018b).
cooling apparatus that included a 0.45 lm cellulose acetate
membrane filter. This sampling, which was conducted more 2.2.2. Characterisation of solids
frequently at the beginning of the experiments due to the The mineralogical composition of the unreacted basalt
faster rate of change expected in the measured parameters, powders and of the solids recovered from the reactors were
decreased the water:rock ratio to approximately 5 by the analysed by powder X-ray diffraction (PXRD, PANalytical

Table 1
Experimental parameters of the basalt-seawater experiments.
Starting materials/g
Experiment T/°C Basalt type Seawater Basalt Water/rock mass ratio Duration/days
BSW04 250 Glassy 275.0 27.50 10 53
BSW05 250 Crystalline 275.0 27.50 10 53
BSW06 290 Glassy 220.0 22.01 10 92
BSW07 290 Crystalline 220.0 22.00 10 93

Table 2
Composition (wt.%) of the basaltic glass and crystalline basalt used as starting materials.
Sample SiO2 Al2O3 FeOc MgO CaO Na2O K2O TiO2 MnO P2O5 Total
a
Glass 48.1 14.6 10.8 9.1 11.8 2.0 0.3 1.6 0.2 0.2 98.7
Crystallineb 47.9 13.4 11.1 10.0 12.2 1.5 0.3 1.6 0.2 0.2 97.9
a
From Oelkers and Gı́slason (2001).
b
From Gudbrandsson et al. (2011).
c
Most iron is Fe2+, Fe2+/Fe3+  10 (Oelkers and Gı́slason, 2001).
24 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

X’Pert PRO, Co-Ka radiation at 40 kV and 40 mA). with previously reported values (Table 3). The column pro-
Briefly, the solid materials were mixed with 20 wt.% of cedure was repeated 2–3 times to achieve a sufficient purity
ZnO as the internal standard, then randomly oriented for Mg analyses, which was confirmed by measurement of
preparations were created using the top loading technique Li, Na, Mg, Al, K, Ca, Ti, V, Mg, Fe, Ni, Sr, Sn and Pb
and subsequently examined in the range 4–85° 2H using a concentrations using a ThermoFisher X-Series ICP-MS.
step size of 0.02° 2H/s with a count time of 1 s/step. Despite the efficiency of this procedure, complete removal
Rietveld-based phase quantification was performed apply- of Ca was not achieved for four samples (BSW04-03,
ing the PANalytical X’Pert HighScore Plus software and BSW05-07, BSW07-03, and BSW07-04), resulting in iso-
the pdf-4 database. For further analysis of the neo- baric interferences during the MC-ICP-MS measurements;
formed clay minerals, the <2 mm size fraction was separated these analyses were corrected as described below.
by centrifugation. Subsequently, oriented mounts were cre- Purified samples were introduced to the MC-ICP-MS in
ated as described in Baldermann et al. (2014), and analysed 3% aqueous HNO3 via a standard injection system (SIS).
in the range 3–30° 2H using a step size of 0.02° 2H/s and a Thirty measurements with a combined integration time of
count time of 2 s/step on a Philips PW 1830 diffractometer 252 s were made per sample. The samples, containing
(Cu-Ka radiation, 40 kV and 30 mA), each under air dried 600 ppb Mg typically yielded 24Mg beam intensities of
conditions, after solvation with ethylene glycol (EG) and 13.5 V. Mass fractionation occurring during the measure-
after heating of the specimens to 550 °C for 1 h. The chem- ments was corrected using the sample-standard bracketing
ical composition and hydration state of the saponite-type technique, and values are reported as
precipitates were estimated from EG-solvated preparations ð 26 Mg= 24 MgÞSample
d26=24
Mg  ð 26 Mg= 24 MgÞ  1 relative to the DSM3 stan-
using the program SybillaÓ developed at ChevronTexaco DSM3

(Baldermann et al., 2013). In addition, the basalt powders dard (Galy et al., 2003). Similar to the method outlined
and reaction products were analysed by scanning electron by Paton et al. (2010), the drift of standard measurements
microscopy (SEM) using a Jeol JSM 6360LV equipped with was modelled using smoothed cubic splines instead of sim-
an electron dispersive spectroscope (EDS). ple linear interpolations for a more accurate representation
of changes in standard measurements based on statistical
2.2.3. Isotopic composition measurements methods. Uncertainties in this drift model and overdisper-
The Mg isotopic compositions of basalt and liquid sam- sion correction were estimated using Monte Carlo simula-
ples were determined by multiple collector inductively cou- tions, so that the given uncertainties were propagated
pled plasma mass spectrometry (MC-ICP-MS) using a from uncertainties in the sample and standard measure-
ThermoFisher Neptune at the National Oceanography ments. All samples were run in duplicate, and their
Centre, Southampton. The solid samples were first dis- weighted means and propagated uncertainties are given in
solved in a mixture of bi-distilled HNO3 and HF at 130 ° Table 4. The long-term 2 standard deviations in d26/24Mg
C, evaporated to dryness, and re-dissolved in double- determinations at the National Oceanography Centre,
distilled HCl. Prior to analysis, Mg was purified using Southampton is 0.07‰, similar to other studies (Pogge
ion-exchange chromatography with a protocol adapted von Strandmann et al., 2011, 2012; Pearce et al., 2012;
from Pogge von Strandmann et al. (2008; 2012). For this, Teng, 2017). The magnitude of 2 standard deviations of
a sample aliquot containing 2000 ng Mg was evaporated the measurements made during this study varied from
to dryness and re-dissolved in 200 mL of 0.8 mol/kg aque- 0.07‰ to 0.24‰. The higher variability is reflected in the
ous HNO3 before loading onto pre-cleaned Bio-Rad uncertainties propagated using the method described
AG50W-X12 (200–400 mesh) resin equilibrated with above.
0.8 mol/kg HNO3 in a Savillex 30 mL column to a height Isobaric interference effects incurred during analysis
of 8.5 cm. Cations other than Mg were eluted using a total from the residual Ca in four samples (BSW04-03,
volume of 51 mL of 0.8 mol/kg aqueous HNO3, followed BSW05-07, BSW07-03, and BSW07-04) were corrected by
by the elution of Mg in 20 mL of 2 mol/kg aqueous assuming that observed deviations of the measured
HNO3. To ensure a 100% yield, this elution protocol was d26/24Mg and d25/24Mg values from the mass dependent
tested using the IAPSO seawater standard, the DSM3 Mg fractionation line (Young et al., 2002) were caused by
48
standard (Galy et al., 2003) as well as the BCR-2 basaltic Ca interference with the 24Mg signal. This enabled the
rock standard and SCo-1 shale standard from the USGS; measured d values to be corrected by back-tracking to the
each showed Mg isotopic compositions in close agreement ‘true’ values that were consistent with mass dependent frac-

Table 3
Measured Mg isotopic composition of standards processed through the column separation protocol in comparison to previously published
values.
Standard Measured d26/24Mg/‰ Literature value/‰ Literature references
IAPSO 0.93 ± 0.20 0.83 ± 0.09 (2SD) Foster et al. (2010); Ling et al. (2011)
0.82 ± 0.07
DSM-3 0.04 ± 0.28 0 (by definition)
0.04 ± 0.07
BCR-2 0.29 ± 0.20 0.31 ± 0.27 (2SD) Pearce et al. (2012)
SCo-1 0.89 ± 0.07 0.92 ± 0.06 (2SD) Li et al. (2010); Ma et al. (2015)
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 25

tionation (see Supplementary Fig. 1). Two of the affected can be explained by the formation of low-NaCl vapour
samples were analysed by MC-ICP-MS after two and three due to the removal of liquid by sampling from the auto-
repeats of the ion-exchange separation, which lead to an claves, leading to an enrichment of solutes in the liquid
increasingly better, but not complete, Ca removal. For both phase (cf. Voigt et al., 2018b for a detailed discussion).
samples, these two individual analyses are consistent with The reduced liquid volume due to this effect is taken into
the same corrected d26/24Mg and d25/24Mg. Furthermore, account in the mass balance calculations presented below,
the analysis after three separation runs indicates a smaller where estimates of the amount of low-NaCl vapour were
amount of 48Ca (Supplementary Fig. 1), consistent with determined using thermodynamic relationships for the
the assumption that the deviations are caused by Ca inter- H2O-NaCl system reported by Driesner (2007) and
ference. Propagation of uncertainties in this inversion was Driesner and Heinreich (2007). To a lesser extent, uptake
performed using Monte Carlo simulations, so that the of H2O due to the formation of hydrated minerals leads
uncertainties given for d26/24Mg and d25/24Mg of the cor- to the same effect.
rected samples include uncertainties of the interference
correction. 3.1.2. Magnesium isotopic compositions
The basaltic glass used in this study had a d26/24Mg com-
3. RESULTS position of 0.28 ± 0.18‰, while the crystalline basalt had
a slightly lower value of 0.38 ± 0.18‰ (uncertainties
3.1. Evolution of the liquid phase composition listed here reflect the 95% confidence interval, correspond-
ing to 2 standard errors). These values are thus identical
3.1.1. General chemical trends within uncertainty, and are in agreement with the value of
The elemental composition of the experimental seawater 0.30 ± 0.03‰ obtained by Wimpenny et al. (2010) for
changed significantly during its reaction with basalt the Stapafell basaltic glass. The seawater used as reactant
(Table 4). The rate of change was higher at 290 °C than had a starting d26/24Mg composition of 0.86 ± 0.07‰,
at 250 °C, and faster reaction progress, as monitored by which is identical within uncertainty to the d26/24Mg value
the temporal evolution of reaction fluid compositions, of 0.83 ± 0.09‰ for seawater recommended by Teng
was also observed for the experiments performed using et al. (2015). Starting from the initial seawater composition,
basaltic glass compared to those using crystalline basalt the liquid evolved towards isotopically lighter compositions
as the reactant, even though the initial surface area of the (Fig. 2, Table 4): In the 250 °C basaltic glass experiment
crystalline material (determined using the BET technique) d26/24Mg initially decreased to 1.02 ± 0.19‰ during the
was approximately 4 times higher. Magnesium was con- first 24 h, then decreased to 3.27 ± 0.87‰ between days
stantly removed from the liquid through basalt-seawater 4 and 7 before the aqueous Mg concentrations became
interaction and reached levels below the detection limit too low for isotopic analysis. In the corresponding crys-
after 7 days in the basaltic glass experiment at 250 °C, while talline basalt experiment, a fluid d26/24Mg value of 1.11
the final aqueous Ca concentrations were higher than the ± 0.13‰ was reached within the first 14 days. The next
initial seawater in both experiments (Fig. 2). Similar to sample, taken after 21 days of total elapsed time, had a
Mg, the SO24 initially present in the seawater was removed
d26/24Mg composition of 0.79 ± 0.14‰, then decreased
from the liquid over time. A pH of 3.7 and 4.9 was attained again to 1.26 ± 0.41‰ at the end of the experiment
in the crystalline basalt experiments at 250 °C and 290 °C, (Fig. 2). At 290 °C, the fluid that was reacted with basaltic
respectively, before increasing to approximately pH 9.5. glass decreased to a d26/24Mg value of 1.07 ± 0.07‰
In contrast, the pH values did not decrease significantly within the first day, while later samples did not contain suf-
in the basaltic glass experiments before they increased to ficient Mg for isotopic analysis. At the same temperature,
similar values as in the crystalline basalt experiments seawater reacted with crystalline basalt showed a slight
(Fig. 2). Significant concentrations of aqueous Fe decrease in d26/24Mg values to 1.01 ± 0.07‰ within
(2.4 mmol/kg and 4.7 mmol/kg at 250 °C and 290 °C, 4 days, followed by a drop to 1.63 ± 0.29‰ and another
respectively, Supplementary Fig. 2) were reached and sus- slow decrease to 1.75 ± 0.08‰ after 35 days.
tained in the experiments with crystalline basalts, while
the liquids had an acidic pH. Measurable quantities of 3.2. Solid alteration products
aqueous silica were present in all experiments from the first
sample (which was taken after 24 h), stabilizing at 10– The basalt powders used as reactants were significantly
15 mmol/kg after decreasing from slightly higher values altered during the experiments. SEM images of the solids
(20 and 25 mmol/kg at 250 °C and 290 °C, respectively). recovered from the reactors after the experiments were ter-
Aluminium concentrations in the liquid remained low minated (Fig. 3) show that the basalt grains were com-
(8 lmol/kg with a standard deviation of 5 lmol/kg), but pletely covered by alteration minerals with honeycomb-
were higher compared to the initial seawater. like structures, which is typical for smectite clays (e.g.,
Aqueous Na and Cl concentrations increased signifi- Fiore et al., 2001). Furthermore, anhydrite was formed dur-
cantly at the end of most experiments: The last two fluid ing all experiments (Voigt et al., 2018b). The massive for-
samples collected from the experiments performed with mation of smectite minerals was confirmed by
basaltic glass at 250 °C had 17% higher Cl concentrations quantitative PXRD analyses (Table 5, Fig. 4). Modelling
compared to seawater, and both experiments at 290 °C of XRD patterns obtained from EG-solvated, oriented clay
showed a similar trend (with an up to 33% increase). This films enabled the smectite clays to be identified as trioctahe-
26
Table 4
Chemical composition of the seawater used as reactant in the experiment and of the liquid samples taken during the experiments. Rows labelled ‘‘Quenched” show compositions of the liquid in the
reactor after terminating the experiment by quenching. Uncertainties (±) designate the 95% confidence interval (2 standard errors) and ’<LOD’ are values below the detection limits, which are
given below the table. Uncertainties for all data are given in Electronic Supplementary Table 1. The major elemental composition was previously reported by Voigt et al. (2018b).
Duration/ pH Fluidmass/g Concentration/mmol/kg d25/24Mg/‰ d26/24Mg/‰
days
Experiment Sample Na Mg Si K Ca Fe Cl SO2
4 Value ± Value ±
Seawater 8.14 474 55 <LOD 11.8 10.3 0.040 597.4 29.1 0.412 0.053 0.859 0.066
BSW04 BSW04-01 1.0 7.71 275 489 17.0 14.4 12.9 16.8 0.27 599.9 0.604 0.54 0.10 1.02 0.19

M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35


250 °C BSW04-02 3.8 8.72 257 491 0.56 13.8 13.6 35.8 0.0069 606.5 0.305 0.53 0.12 1.02 0.22
basaltic glass BSW04-03 6.8 9.09 235 493 0.106 13.4 13.7 39.7 0.0008 619.0 0.287 1.67 0.44a 3.27 0.87a
BSW04-04 13.9 9.48 212 500 <LOD 10.5 13.8 43.6 <LOD 620.4 0.295
BSW04-05 21.0 9.58 190 504 <LOD 9.8 13.8 45.8 <LOD 624.5 0.289
BSW04-06 33.9 9.65 169 575 <LOD 9.2 15.9 42.8 <LOD 684.6 0.345
BSW04-07 52.9 9.52 148 601 <LOD 10.5 16.9 36.9 <LOD 704.1 0.350
Quenched 513 <LOD 4.25 9.6 46.9 <LOD 621.9 4.00
BSW05 BSW05-01 1.0 3.74 275 484 46.7 18.2 14.7 1.28 1.37 608.0 14.85 0.42 0.06 0.80 0.12
250 °C BSW05-02 3.8 4.01 254 479 42.9 20.4 15.2 1.43 2.0 603.8 12.45 0.44 0.06 0.85 0.13
crystalline BSW05-03 6.8 4.83 232 490 40.4 20.6 15.8 1.63 2.4 610.8 10.15 0.48 0.06 0.92 0.13
basalt BSW05-04 14.0 5.89 212 480 20.9 17.1 16.1 10.3 1.43 597.5 0.902 0.58 0.06 1.11 0.13
BSW05-05 21.0 9.03 192 489 0.84 6.0 16.7 35.7 0.014 608.3 0.291 0.41 0.07 0.79 0.14
BSW05-06 33.9 9.01 172 494 0.31 10.4 16.7 37.7 0.007 620.1 0.263 0.47 0.07 0.89 0.14
BSW05-07 52.9 9.08 151 502 0.172 9.9 17.1 40.6 0.004 628.9 0.254 0.64 0.21a 1.26 0.41a
Quenched 482 <LOD 3.97 16.1 44.5 <LOD 622.0 4.00
BSW06 BSW06-01 0.8 8.36 220 524 6.9 23.7 12.0 28.6 <LOD 636.5 0.181 0.588 0.063 1.069 0.066
290 °C BSW06-02 3.0 9.20 200 541 0.078 19.7 12.3 42.4 <LOD 643.5 0.124
basaltic glass BSW06-03 7.9 9.40 180 545 0.0154 13.5 12.6 45.4 <LOD 654.6 0.120
BSW06-04 16.0 9.41 160 550 0.0080 15.7 13.5 46.3 <LOD 667.1 0.123
BSW06-05 33.8 9.50 141 561 0.0062 13.9 14.2 44.9 <LOD 689.6 0.122
BSW06-06 54.9 9.56 122 582 0.0105 12.4 18.9 62.4 <LOD 736.3 0.141
BSW06-07 92.0 9.50 104 608 0.0200 10.6 23.9 78.8 <LOD 806.6 0.132
BSW07 BSW07-01 0.9 4.93 219 512 30.6 25.8 14.3 2.25 4.7 631.3 3.951 0.558 0.062 1.066 0.070
290 °C BSW07-02 3.8 8.40 201 516 5.0 9.8 15.5 29.9 0.50 648.0 0.175 0.531 0.061 1.011 0.066
crystalline BSW07-03 8.7 8.92 183 515 0.55 14.6 15.6 38.0 0.044 656.7 0.132 0.83 0.15 1.63 0.29
basalt BSW07-04 16.9 9.01 165 511 0.41 15.6 15.8 39.3 0.033 660.2 0.124 0.87 0.14 1.71 0.26
BSW07-05 34.6 9.03 148 528 0.29 16.6 16.4 41.7 0.032 678.0 0.062 0.888 0.063 1.746 0.077
BSW07-06 55.7 9.13 128 543 0.201 16.3 17.0 43.8 <LOD 700.5 0.057
BSW07-07 92.8 9.22 110 589 0.127 14.8 18.4 48.6 <LOD 754.2 0.113
Detection limits (LOD) in mmol/kg were approximately: Mg: 0.026; Si: 0.018; Fe (BSW04 and BSW05): 0.0001; Fe (BSW06 and BSW07): 0.02.
a
These uncertainties include the extrapolation of 48Ca interference effects.
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 27

26/24
Fig. 2. Aqueous concentrations of Ca, Mg, SO24 as well as d Mg and pH of seawater reacting with basaltic glass and crystalline basalt at
temperatures of 250 °C (left side) and 290 °C (right side). The errors bars show the 95% confidence intervals, which are smaller than the
symbol size where no error bars are shown.

dral ferrous saponite with Na ions occupying mainly the abundant precipitation of ferrous saponite and minor anhy-
exchangeable interlayer sites (Fig. 4). The ferrous saponite drite, which removed Mg as well as Na, OH, Al, Si, Fe, Ca,
precipitated in this study contains approximately 0.46– SO24 and Sr from the fluid (Voigt et al., 2018b). A number
0.67 atoms per formula unit (a.p.f.u.) octahedral Fe in the of studies have previously concluded that the evolution of
experiments with crystalline basalt, and from 0.71 to 0.98 fluid pH in such systems is controlled by the relative impor-
a.p.f.u. octahedral Fe in the basaltic glass experiments tance of two competing processes: (1) The uptake of alka-
(Table 5; Supplementary Fig. 3). Furthermore, PXRD data linity (i.e. OH groups) along with Mg, Si and Fe by
suggest that minor analcime/heulandite type zeolite formed secondary hydrous minerals, which drives a drop in pH
during the two basaltic glass experiments. At 290 °C, pla- as seen in the beginning of the experiments with crystalline
gioclase (close to albitic composition) and minor propor- basalt; and (2) leaching of Ca and other cations from the
tions of mixed-layer chlorite-smectite minerals were basalt causing pH and aqueous Ca concentrations to
formed in the experiment with basaltic glass according to increase after Mg removal was near complete (Bischoff
the PXRD analyses (Table 5, Fig. 4). and Dickson, 1975; Seyfried and Bischoff, 1977; Seyfried
and Bischoff, 1979; Seyfried and Bischoff, 1981; Seyfried
4. DISCUSSION and Mottl, 1982). The stronger increase in pH observed
in this study compared to previous experiments and marine
4.1. Evolution of liquid chemistry and alteration reactions high-temperature hydrothermal fluids (Fig. 2; Seyfried and
Bischoff, 1979; Seyfried and Bischoff, 1981; German and
The reaction of basalt with seawater under hydrother- Seyfried, 2014) is possibly related to the uptake of HCl by
mal conditions in our experiments caused the rapid and the gas phase enabled by the high salinity, and/or the
28 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

(a) (b)

Anhydrite

50 μm 50 μm

(c) (d)

Smectite

Smectite
2 μm 1 μm

Fig. 3. Scanning electron microscopy images of (a) the basaltic glass used as starting material and (b)–(d) solids recovered from the
experiment reacting basaltic glass with seawater at 250 °C, showing a layer of smectite clay minerals covering the surface of the altered basalt
as well as minor precipitated anhydrite grains. Images of the crystalline basalt are shown in Voigt et al. (2018b).

Table 5
Proportions (wt.%) of phases identified by quantitative XRD measurements in the basaltic glass and crystalline basalt used as reactants in the
experiments, as well as in the solids recovered from the reactors after the experiments were terminated. The last row shows the total octahedral
Fe contents of the smectite, estimated using the position of the d060 reflection and the relationship found by Baldermann et al. (2014).
Basaltic glass BSW04 BSW06 Crystalline basalt BSW05 BSW07
Amorphous 95 24 15 – – –
Quartz 2 <1 <1 <1 <1 <1
Olivine 3 <1 2 9 <1 <1
Anhydrite – <1 1 – 4 2
Zeolite – 8 5 – – –
Plagioclase – – 13 43 53 42
Pyroxene – – – 47 25 23
Chlorite-smectite – – 8 – – –
Fe-rich saponite – 66 55 – 18 33
VI
Fetot/a.p.f.u. 0.71 0.98 0.46 0.67

presence of small amounts of NH3 introduced during the increasing more Fe-rich saponite during subsequent min-
pH adjustment of the seawater used in the experiments eral growth.
(see Voigt et al., 2018b). The low initial pH in the crystalline
basalt experiments also explains the prolonged period of Fe 4.2. Magnesium isotope fractionation
enrichment in the liquid (released from basalt), after which
it was incorporated into the saponite clays as indicated by The changes in aqueous d26/24Mg during the experi-
their high octahedral Fe contents (cf. Section 3.2) and the ments likely reflect the combined effects of isotopic fraction-
absence of secondary Fe-hydroxide minerals. Fluid enrich- ation during saponite precipitation and the Mg release from
ment in Fe is also consistent with trends observed in natural basalt dissolution. Consequently, an assessment of the
marine hydrothermal fluids (e.g., German and Seyfried, extent of isotope fractionation into the alteration products
2014). Furthermore, the correlation between the total mass requires an estimate of the amount of Mg released from
of precipitated smectite and the estimated structural Fe basalt dissolution, which is later scavenged by clay mineral
content of saponite formed in experiments with basaltic formation. Here, we estimate these values from mass bal-
glass and crystalline basalt (Table 5) suggests that a Fe- ance relationships between aqueous and solid phase Mg,
poor saponite was initially precipitated, followed by an assuming that all Mg released from basalt dissolution is
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 29

a)11000 Basalt glass


BSW04
b) Crystalline basalt
BSW05
BSW06 BSW07
Aug
9000 Pl
Qz
Pl
Diffracted Intensity (c.p.s.)

Ol
Qz
Ol Ol
7000
Qz
Pl Pl
Ol Ol Ol Aug Ol
Pl

5000 Zeo
Zeo
Zeo
Zeo

Zeo
Pl
Zeo

3000
Zeo

Chl-Smc Pl
001 Zeo Pl 001
Fe-Sap± Fe-Sap
Chl-Smc 02,11 004
Anh 13,20 15,24,31 060 02,11 Anh 13,20 15,24,31 060
1000
5 15 25 35 45 55 65 75 85 5 15 25 35 45 55 65 75 85
°2-Theta (Co-Kα) °2-Theta (Co-Kα)

Fig. 4. PXRD patterns of starting materials and reacted solids from experiments with basaltic glass (a) and crystalline basalt (b). Note the
presence of amorphous phases in all samples with basaltic glass. The hkl-reflections correspond mainly to ferrous saponite. Zincite (20 wt.%;
bars) was used as an internal standard. Fe-sap = ferrous saponite; Chl-Smc = chlorite-smectite; Zeo = analcime/heulandite; Pl = plagioclase
± albite; Anh = anhydrite; Qz = quartz; Ol = olivine; Aug = augite.

precipitated into a single alteration phase (i.e. saponite). It sible to reproduce the Mg isotope trends observed in our
is also assumed that no significant Mg is taken up by the experiments without including isotope fractionation during
low-NaCl vapour phase that is present at the end of the basalt dissolution, and preliminary model runs that did
experiments, as the vapour pressure of Mg is significantly include constant isotopic fractionation during the release
lower than that of Na (Lide, 2003); thus Mg is only of Mg from basalt dissolution did not significantly improve
expected to be present in trace quantities in the vapour 26=24
the model fit (resulting in aSmectiteLiquid values similar to
phase under the experimental conditions used in this study
those presented below). In contrast, it was found that differ-
(Driesner, 2007; Driesner and Heinrich, 2007). The amount
ent fractionation factors for the two experimental tempera-
of Mg released from basalt dissolution was estimated from
mass balancing the components Mg, Ca and SO4. We tures applied in this study were necessary to explain the
assume that Ca and SO4 are only removed from the liquid isotopic trends with this mass balance model.
by precipitation into anhydrite. The latter assumption is The results, shown in Fig. 5c–d, reproduce the major
likely not entirely satisfied, especially during the later stages trends seen in the experimental d26/24Mg data: The
of the experiments, when Ca could be incorporated into one d26/24Mg composition of the fluid decreased slowly during
or more of the other observed alteration phases (saponite, the early stages of the experiments following the preferen-
zeolites, and albitic plagioclase). Therefore, the calculated tial incorporation of heavy Mg into smectite
26=24
amounts of Mg released from basalt dissolution (symbols (eSmectiteLiquid > 0). The short-lived increase in liquid
in Fig. 5a,b) and Mg incorporation into smectite are likely d 26/24
Mg values observed in the 250 °C crystalline basalt
minimum estimates. experiment may be attributed to the release of isotopically
Based on the mass balance constraints described above, heavy Mg from basalt dissolution (as can be seen from
the evolution of d26/24Mg during the experiments was mod- the mass balance and strongly increasing fluid Ca concen-
elled using an incremental isotope mass balance model, trations, Figs. 1 and 3), which likely results from an
allowing for isotopic fractionation during the precipitation increase in Al removal from the fluid into secondary miner-
of a single Mg-incorporating mineral (saponite). A constant als when pH increases. Conversely, the strong decrease of
isotopic fractionation factor describing irreversible isotopic
d26/24Mg in the liquid towards the end of the experiments
fractionation during smectite precipitation (i.e. the smectite
is caused by the irreversible (Rayleigh-type) nature of the
does not react with the liquid after precipitation) is assumed
Mg fractionation into smectite.
in this model, and is defined as
The best estimate and uncertainty (i.e. analytical uncer-
26=24 d26 MgSmectite =d24 MgSmectite tainties propagated through the model, 95% confidence
aSmectiteLiquid  ð4:1Þ
Liquid = MgLiquid
26=24
26 Mg 24
intervals) for eSmectiteLiquid (defined as e26=24 ¼ ða26=24  1Þ)
(see Appendix A for model details). The resulting model is 0:422  0:065‰ at 250 °C and 0:346  0:016‰ at 290 °
26=24
can be interpreted as a Rayleigh type fractionation process, C. These values are similar to the eBruciteLiquid ¼ 0:5 value
whereby the changes in d26/24Mg are primarily driven by obtained experimentally by Wimpenny et al. (2014) for
smectite precipitation from a limited volume of liquid. brucite at 80 °C, which is used as an analogue to describe
The possibility of coincident Mg isotopic fractionation Mg incorporation into clay minerals, and are qualitatively
occurring during basaltic glass dissolution was suggested in agreement with a decrease in the magnitude of
by experiments of Wimpenny et al. (2010) and for Si iso- isotopic fractionation with increasing temperature. They
topes by data of Ziegler et al. (2005). However, it was pos- are also consistent with the fractionation factor of
30 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

Fig. 5. Estimated mass of Mg released from basalt dissolution and precipitated into alteration minerals, as calculated by mass balance (a,b,
see text for details), (c,d) modelled aqueous d26/24Mg during the experiments reacting basalt with seawater (dashed curves), (e,f) measured Mg
concentrations in the liquid and (g,h) relative abundance (mol%) of the Mg2+ and MgSO4 species in the fluid samples as calculated with
PHREEQC. The errors bars and shaded areas show the 95% confidence intervals, which are smaller than the symbol size where no error bars
are shown.

26=24 enriched Mg sink (likely clay minerals) relative to the


eClayLiquid ¼ 0:59  0:14‰ determined during synthetic
lizardite and kerolite formation at 250 °C by Ryu et al. pore-fluid. Our results also support the conclusions drawn
(2016), as well as with the behaviour of Mg during saprolite by Huang et al. (2018), who studied the Mg isotope compo-
formation in the context of diabase weathering, as Teng sitions of oceanic crustal rocks that have been subjected to
et al. (2010) suggested fractionation factors of around low-temperature alteration, and inferred a preferential
26=24 incorporation of heavy Mg isotopes in sections where sapo-
eSaproliteFluid  0:4‰ for smectite-rich saprolite at Earth sur-
nite formation occurred. Furthermore, silicate soil samples
face temperatures. The preferential enrichment of smectite from the Southern Tibetan Plateau show higher d26/24Mg
in heavier Mg isotopes is also in agreement with the deep- values compared to associated silicate host rocks, a differ-
sea pore-fluid data of Higgins and Schrag (2010), who ence that can be attributed to silicate rock weathering
observed a decrease in pore-fluid d26/24Mg with depth at (Tipper et al., 2006a). Conversely, the direction of Mg iso-
some sites, indicative of the presence of an isotopically tope fractionation observed in this study is opposite to the
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 31

results of Pogge von Strandmann et al. (2008), who sug- where . . . reflects the contributions of other fluxes on the
gested that secondary minerals such as talc and chlorite marine Mg budget (e.g. Tipper et al., 2006b; Pogge von
preferentially uptake isotopically light Mg from high pH Strandmann et al., 2014; Higgins and Schrag, 2015).
fluids. Assuming that irreversible isotopic fractionation occurs in
Variations in Mg isotope fractionation factors during a closed system, it can be shown (Appendix B) that for a
clay mineral formation can be explained by a number of given fraction f of Mg removal from the seawater, the iso-
parameters, including temperature, the mineral composi- topic evolution can be described by
tion, reaction kinetics, and fluid speciation. For example,   ðð1  f Þe  1Þðf  1Þ
Schott et al. (2016) showed that the presence of organic D26=24 Mg  ð4:3Þ
ClaysSW f
complexes in seawater can have significant effects on Mg
aqueous speciation in seawater and thus Mg isotopic frac- This function has its maximum of e at f = 0 and is shown
26=24
tionation during mineral precipitation. The speciation of in Fig. 6a for eSmectiteLiquid ¼ 0:35‰, which is the mean of
aqueous Mg in our experiments, as calculated with our experimental best fit weighted by the inverse of their
PHREEQC (Parkhurst and Appelo, 2013) together with variance. Since FMg / f, the effect of Mg isotope fraction
the CarbFix database (Voigt et al., 2018a) at the experimen- in these clays is proportional to
tal temperatures (Fig. 5g–h), is dominated by the MgCl+
complex, representing between 66 and 90 mol% of the total EClays / ðð1  f Þe  1Þðf  1Þ ð4:4Þ
aqueous Mg in the initial seawater, while the Mg2+ complex illustrating that the effect of clay precipitation on the mar-
is estimated to have an abundance of 10–20 mol%, and the ine budget is largest at f  1  1=e  0:63 for a constant e,
initially present aqueous MgSO4 (up to 16 mol%) decreased and is zero at both no clay mineral precipitation and com-
with the rapid SO4 removal from seawater. The degree to plete Mg removal (Fig. 6b).
which the fractionation factors determined in this study This result shows that the impact of Mg isotopic frac-
might be affected by aqueous Mg speciation remains unre- tionation during smectite formation in high-temperature
solved at present. marine hydrothermal systems on the marine Mg isotopic
budget may be low, as Mg is almost quantitatively removed
4.3. Consequences for the marine Mg isotopic budget from the seawater during hydrothermal reactions (e.g.,
Mottl and Holland, 1978; Elderfield et al., 1999; Coggon
Recent studies of the marine Mg isotopic budget suggest et al., 2004; German and Seyfried, 2014). Using the mean
that variations in carbonate mineral formation (such as
dolomitization) and riverine inputs have exerted a major
influence on the composition of seawater through time
(Pogge von Strandmann et al., 2014; Higgins and Schrag,
2015), and that a decrease in seawater d26/24Mg over the
past 15 Ma may have been driven by a decrease in the
extent of dolomite precipitation (Pogge von Strandmann
et al., 2014). Although the influence of Mg-clay formation
in marine hydrothermal environments and in marine sedi-
ments on the global marine budget has been proposed by
Higgins and Schrag (2015) and Dunlea et al. (2017), the
magnitudes of these effects are poorly constrained at pre-
sent, and are expected to vary as a function of the degree
of Mg removal in these settings. Our experimental results
provide the first evidence for Mg isotope fractionation dur-
ing clay formation in intermediate temperature hydrother-
mal settings, and imply that the return of isotopically
light fluids that have undergone partial Mg removal into
clay minerals to seawater might significantly impact
d26/24Mg mass balance.
The potential effect of Mg isotope fractionation in these
systems on the marine Mg budget (here called EClays) is a Fig. 6. Plot of the difference between the Mg isotopic composition
combination of: (i) the isotopic composition (i.e. the of the reacting seawater (SW) and its initial composition,
 
d26/24Mg value) of the precipitated clay minerals relative D26=24 Mg , blue curve, and between the accumulated
ReactedSWSW  
to seawater; and (ii) the flux of Mg between these systems
clays and the initial SW, D26=24 Mg , orange curve, during
F Clays
Mg , since ClaysSW
irreversible isotopic fractionation during clays precipitation as
  function of reaction progress, f ¼ mP =mR;0 . Both curves are plotted
d d26=24 Mg     
 cSW for e ¼ 0:35‰, the mean (weighed by the inverse of their variance)
Mg ¼ d26=24 Mg  d26=24 Mg  F Clays
Mg þ . . .
SW
dt Clays SW of our experimental best fit values. The latter value, scaled by the
 
 D26=24 Mg  F Clays fraction of Mg in the product, f, is shown as well by the green curve
Mg þ . . . ð4:2Þ
Clays - SW
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} in (b) to qualitatively show the effect of Mg isotope fractionation in
EClays such a system on the marine budget.).
32 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

of our best fit fraction factor, the isotopic difference 0:346  0:016‰ at 290 °C indicate that isotopically heavy
between seawater 
and clays due
 to irreversible precipitation Mg is preferentially incorporated into the smectite, explain-
is approximately D26=24 Mg  0:14‰ for an 80% ing the drop in liquid d26/24Mg when aqueous Mg concen-
ClaysSW trations approach zero.
removal of Mg from seawater (see Fig. 6), which is assumed 26=24
The eSmectiteLiquid values determined in this study are in
to be the minimum value for high temperature hydrother-
agreement with independent field and experimental evi-
mal systems, and it often approaches 100% (Mottl and
dence for similar silicate minerals (Tipper et al., 2006a;
Wheat, 1994; German and Seyfried, 2014).
Higgins and Schrag, 2010; Teng et al., 2010; Wimpenny
Conversely, for low temperature alteration, Mg removal
et al., 2014), and provide the first constraint on the
is less complete and reaches down to f = 0 (Fisher and
d26/24Mg composition of smectite forming under hydrother-
Wheat, 2010). In their marine Mg budget model, Higgins
mal conditions. The observation that smectite preferentially
and Schrag (2015) adopted a value of 0:7‰ for the isotopic
incorporates isotopically heavy Mg is opposite to the trend
difference between seawater and Mg-clays,
  seen in carbonates (Galy et al., 2002; Tipper et al., 2006b;
D26=24 Mg , while Huang et al. (2018) adopted a Higgins and Schrag, 2010; Teng et al., 2010; Pearce et al.,
Clays-SW
value of 0:83‰ based on the average composition of altered 2012; Mavromatis et al., 2013; Huang et al., 2018), thus
oceanic crust at Site 801. The fractionation factors deter- highlights the potential utility of Mg isotopes in fluids as
mined under our experimental conditions are tracers of the formation of these different mineral types
  (e.g., Oelkers et al., 2018). Furthermore, the magnitude of
D26=24 Mg < 0:350‰ (Fig. 6), which is slightly
Clays-SW Mg fractionation observed during clay formation in this
smaller than the values adopted by Higgins and Schrag study is in good agreement with the previously estimated
(2015) and Huang et al. (2018). However, Mg isotopic frac- effects of Mg isotopic fractionation in low-temperature
tionation is expected to more pronounced at lower temper- marine hydrothermal systems on the marine Mg budget
atures (e.g., Teng, 2017), thus these values are qualitatively (Higgins and Schrag, 2015; Huang et al., 2018), significantly
in agreement. reducing uncertainties in the modelling of the Mg isotopic
In comparison, the influence of carbonate precipitation compositional evolution in the marine Mg cycle.
on the Mg isotopic budget of the oceans is approximately
an order of magnitude higher due to its larger isotopic frac- Declaration of Competing Interest
26=24
tionation factor of eCarbonateLiquid  2‰ (Higgins and
Schrag, 2015), even considering a higher Mg flux at The authors declare that they have no known competing
mid-ocean ridges compared to that of marine carbonate financial interests or personal relationships that could have
formation. Nevertheless, both carbonate formation and appeared to influence the work reported in this paper.
clay precipitation with incomplete Mg-removal from the
seawater are crucial when constraining the global marine ACKNOWLEDGEMENTS
Mg cycle using isotopes, as suggested by Higgins and
Schrag (2015) and Huang et al. (2018). This research was supported by the CO2-REACT Marie Curie
Actions Initial Training Network which was funded by the Euro-
pean Union’s Seventh Framework Programme under grant agree-
5. CONCLUSIONS ment no 317235. The authors would like to thank Frederic
Candaudap, Aurélie Marquet, Alain Castillo, Pascal Gisquet,
The evolution of aqueous d26/24Mg was followed during Thierry Aigouy, Stephanie Mounic, Mathieu Benoit, Carole Caus-
basaltic glass and crystalline basalt-seawater interaction serand, Manuel Henry, Jonathan Prunier, Matt Cooper, and Agnes
experiments at temperatures of 250 °C and 290 °C under Michalik for help with analyses and technical assistance. We would
saturated vapour pressure. The solid alteration phases also like to thank Deirdre Clark, Jan Přikryl, and Sigurður Reynir
recovered after termination of the experiments mainly con- Gı́slason for providing the Stapafell basalt, as well as Catherine
sisted of ubiquitous saponite (a trioctahedral Mg-smectite) Jeandel for providing the seawater. Finally, we thank Damon Tea-
and minor anhydrite. Trends in the chemical evolution of gle for his editorial support plus Ed Tipper and two anonymous
reviewers for their insightful and helpful comments.
the fluid composition sampled during the experiments were
similar to those observed at marine hydrothermal systems.
During the experiments, Mg was released into the liquid
phase from basalt dissolution, while the precipitation of APPENDIX A. DETAILS OF THE MASS BALANCE
alteration minerals such as smectite removed Mg from the MODEL CALCULATIONS
liquid, leading to a steady decrease in aqueous Mg concen-
trations in all experiments. The Mg fractionation factors for the formation of smec-
A model of Mg isotopic evolution during the experi- tite from the aqueous fluids were generated from the solu-
ments, which assumes irreversible (Rayleigh-type) fraction- tions to a set of simulation differential equations.
ation during precipitation into smectite concurrent with Mg Solutions to these differential equation system were calcu-
release from basalt dissolution, can fully explain the trends lated numerically with the NDSolve function of Wolfram
observed in the experimentally determined fluid d26/24Mg Mathematica, and the maximum likelihood a26=24 SmectiteLiquid
compositions. The retrieved fractionation factors of (expressed here as e26=24 ¼ ða26=24  1Þ) are calculated based
26=24
eSmectiteLiquid ¼ 0:422  0:065‰ at 250 °C and on the minimum inverse variance weighted squared errors
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 33

using the Nelder-Mead method (Nelder and Mead, 1965). before the reaction, respectively. The composition of the
The corresponding uncertainties are propagated using reactant evolves according to
Monte Carlo simulations where the 95% confidence inter-  
RR
vals are calculated from the quantiles. As mass balance con- ln  e lnð1  f Þ
R0R
straints require inclusion of 25Mg into the equations, a25=24
are calculated internally based on a26=24 and the kinetic where RR and R0R designate the isotopic ratios of the reac-
mass-dependent fractionation law (Young et al., 2002), tant after reaction progress f and at the beginning, respec-
which was used as opposed to the equilibrium mass frac- tively, as shown by Bigeleisen and Wolfsberg (1957).
tionation law due to the relatively fast kinetics of the Substituting RX ¼ ðdX þ 1ÞRStandard and solving for
involved processes. However, due to the small difference dR  dR;0 ; the change in isotopic composition of the reac-
between the two fractionation laws, the choice does not sig- tant, gives
nificantly affect the model results. The mean covariance
dR  dR;0 ¼ ð1 þ dR;0 Þð1  f Þe  1  ð1  f Þe  1
between measured d25/24Mg and d26/24Mg, 0.81, was
included in the uncertainty propagation calculations. To Conversely, Bothner-By and Bigeleisen (1951) showed
account for the generation of low-NaCl vapour due to sam- that the isotopic composition of the total accumulated pro-
pling from the reactors towards the ends of the experi- duct follows the relationship
ments, all mass balance calculations account for the  
R0  RP;Total f
vapour volume estimated using the thermodynamics of ln 1 þ R 0  e lnð1  f Þ
the H2O-NaCl system (Driesner, 2007; Driesner and RR 1f
Heinrich, 2007). where RP;Total designates the isotopic ratios of the accumu-
The evolution of Mg dissolution and precipitation dur- lated product which can be solved in a similar way to give
ing the experiments (symbols in Fig. 5a–b) were estimated the difference dP;Total  dR;0 as
using the Steffen interpolation method (Steffen, 1990) to
establish a smooth trend while preserving the monotonic- ð1 þ dR;0 Þðð1  f Þe  1Þðf  1Þ
dP;Total  dR;0 ¼
ity. The elemental mass balance model furthermore allows f
estimation of molar ratios of elements incorporated into ðð1  f Þe  1Þðf  1Þ

secondary phases during the experiment based on the evo- f
lution of fluid composition and the inferred amounts of
basalt dissolution. We calculated the molar ratio of (Si
+ Al)/(Mg + Fe) taken up into secondary minerals, which APPENDIX C. SUPPLEMENTARY MATERIAL
is 4/3  1.33 in case of saponite, lower than the mean
(weighed by the inverse of the variance) of the corre- Supplementary data to this article can be found online at
sponding value in our experiments of 2.37 ± 0.87. How- https://doi.org/10.1016/j.gca.2019.12.026.
ever, the uncertainty of this parameter is relatively high,
as can be seen in Supplementary Fig. 4, showing its evo- REFERENCES
lution during the experiments. The retrieved (Si + Al)/
Baldermann A., Dohrmann R., Kaufhold S., Nickel C., Letofsky-
(Mg + Fe) values that are greater than expected for sapo-
Papst I. and Dietzel M. (2014) The Fe-Mg-saponite solid
nite can be explained by (i) uptake of excess Si and Al solution series a hydrothermal synthesis study. Clay Miner. 49,
into other phases (such as zeolites, as confirmed in 391–415.
BSW-04 and BSW-06 by XRD measurements), (ii) the Baldermann A., Grießbacher A. C., Baldermann C., Purgstaller B.,
fact that calculated amounts of Mg release from basalt Letofsky-Papst I., Kaufhold S. and Dietzel M. (2018) Removal
dissolution and thus Mg incorporation into smectite are of barium, cobalt, strontium, and zinc from solution by natural
likely minimum estimates (see Section 4.2), or a combina- and synthetic allophane adsorbents. Geosciences 8, 309.
tion of the two. Baldermann A., Warr L. N., Grathoff G. H. and Dietzel M. (2013)
The Rate and Mechanism of Deep-Sea Glauconite Formation
APPENDIX B at the Ivory Coast - Ghana Marginal Ridge.
Bigeleisen J. and Wolfsberg M. (1957) Theoretical and experimen-
tal aspects of isotope effects in chemical kinetics. In Advances in
Irreversible isotopic fractionation in a closed system Chemical Physics (eds. I. Prigogine and P. Debye). John Wiley
& Sons, Inc. pp. 15–76.
Isotopic fractionation occurring during a reaction where Bischoff J. L. and Dickson F. W. (1975) Seawater-basalt interac-
the product is immediately isolated from the reactant can tion at 200°C and 500 bars: Implications for origin of sea-floor
be described by heavy-metal deposits and regulation of seawater chemistry.
Earth Planet. Sci. Lett. 25, 385–397.
RP;i
aProductReactant  ðe þ 1Þ  : Bothner-By A. A. and Bigeleisen J. (1951) Fractionation of the
RR carbon isotopes in decarboxylation reactions. III. The relative
where RR and RP;i designate the isotopic ratios of the reac- rates of decomposition of carboxyl-C12 and -C13 mesitoic acids.
tant and of an increment of the product during the reaction, J. Chem. Phys. 19, 755–759.
Brunauer S., Emmett P. H. and Teller E. (1938) Adsorption of
respectively. The reaction progress is commonly expressed
gases in multimolecular layers. J. Am. Chem. Soc. 60, 309–319.
as fractional yield, defined as f  mP =mR;0 , where mP and
Coggon R. M., Teagle D. A. H., Cooper M. J. and Vanko D. A.
mR;0 are the masses of the product and of the reactant (2004) Linking basement carbonate vein compositions to
34 M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35

porewater geochemistry across the eastern flank of the Juan de clays, seawater Mg/Ca, and climate. Earth Planet. Sci. Lett.
Fuca Ridge, ODP Leg 168. Earth Planet. Sci. Lett. 219, 111– 416, 73–81.
128. Horita J., Zimmermann H. and Holland H. D. (2002) Chemical
Coogan L. A. and Dosso S. (2012) An internally consistent, evolution of seawater during the Phanerozoic: implications
probabilistic, determination of ridge-axis hydrothermal fluxes from the record of marine evaporites. Geochim. Cosmochim.
from basalt-hosted systems. Earth Planet. Sci. Lett. 323–324, Acta 66, 3733–3756.
92–101. Huang K.-J., Teng F.-Z., Plank T., Staudigel H., Hu Y. and Bao
Driesner T. (2007) The system H2O–NaCl. Part II: Correlations for Z.-Y. (2018) Magnesium isotopic composition of altered
molar volume, enthalpy, and isobaric heat capacity from 0 to oceanic crust and the global Mg cycle. Geochim. Cosmochim.
1000 °C, 1 to 5000 bar, and 0 to 1 XNaCl. Geochim. Cosmochim. Acta 238, 357–373.
Acta 71, 4902–4919. Li W., Beard B. L., Li C. and Johnson C. M. (2014) Magnesium
Driesner T. and Heinrich C. A. (2007) The system H2O–NaCl. Part isotope fractionation between brucite [Mg(OH)2] and Mg
I: Correlation formulae for phase relations in temperature– aqueous species: implications for silicate weathering and
pressure–composition space from 0 to 1000 °C, 0 to 5000 bar, biogeochemical processes. Earth Planet. Sci. Lett. 394, 82–93.
and 0 to 1 XNaCl. Geochim. Cosmochim. Acta 71, 4880–4901. Li W.-Y., Teng F.-Z., Ke S., Rudnick R. L., Gao S., Wu F.-Y. and
Dunlea A. G., Murray R. W., Ramos D. P. S. and Higgins J. A. Chappell B. W. (2010) Heterogeneous magnesium isotopic
(2017) Cenozoic global cooling and increased seawater Mg/Ca composition of the upper continental crust. Geochim. Cos-
via reduced reverse weathering. Nat. Commun. 8, 1–7. mochim. Acta 74, 6867–6884.
Elderfield H. and Schultz A. (1996) Mid-ocean ridge hydrothermal Lide, D. R. (ed.) (2003) CRC Handbook of Chemistry and Physics,
fluxes and the chemical composition of the ocean. Annu. Rev. 84th ed. CRC Press, Boca Raton, Florida.
Earth Planet. Sci. 24, 191–224. Ling M.-X., Sedaghatpour F., Teng F.-Z., Hays P. D., Strauss J.
Elderfield H., Wheat C. G., Mottl M. J., Monnin C. and Spiro B. and Sun W. (2011) Homogeneous magnesium isotopic compo-
(1999) Fluid and geochemical transport through oceanic crust: sition of seawater: an excellent geostandard for Mg isotope
a transect across the eastern flank of the Juan de Fuca Ridge. analysis. Rapid Commun. Mass Spectrom. 25, 2828–2836.
Earth Planet. Sci. Lett. 172, 151–165. Ma L., Teng F.-Z., Jin L., Ke S., Yang W., Gu H.-O. and Brantley
Fantle M. S. and DePaolo D. J. (2006) Sr isotopes and pore fluid S. L. (2015) Magnesium isotope fractionation during shale
chemistry in carbonate sediment of the Ontong Java Plateau: weathering in the Shale Hills Critical Zone Observatory:
Calcite recrystallization rates and evidence for a rapid rise in accumulation of light Mg isotopes in soils by clay mineral
seawater Mg over the last 10 million years. Geochim. Cos- transformation. Chem. Geol. 397, 37–50.
mochim. Acta 70, 3883–3904. Mavromatis V., Gautier Q., Bosc O. and Schott J. (2013) Kinetics
Fiore S., Huertas F. J., Huertas F. and Linares J. (2001) Smectite of Mg partition and Mg stable isotope fractionation during its
formation in rhyolitic obsidian as inferred by microscopic incorporation in calcite. Geochim. Cosmochim. Acta 114, 188–
(SEM-TEM-AEM) investigation. Clay Miner. 36, 489–500. 203.
Fisher A. T. and Wheat C. G. (2010) Seamounts as conduits for Mottl M. J. and Holland H. D. (1978) Chemical exchange during
massive fluid, heat, and solute fluxes on ridge flanks. Oceanog- hydrothermal alteration of basalt by seawater—I. Experimental
raphy 23, 74–87. results for major and minor components of seawater. Geochim.
Foster G. L., Pogge von Strandmann P. a. E. and Rae J. W. B. Cosmochim. Acta 42, 1103–1115.
(2010) Boron and magnesium isotopic composition of seawater. Mottl M. J., Holland H. D. and Corr R. F. (1979) Chemical
Geochem. Geophys. Geosystems 11, Q08015. exchange during hydrothermal alteration of basalt by seawa-
Galy A., Bar-Matthews M., Halicz L. and O’Nions R. K. (2002) ter—II. Experimental results for Fe, Mn, and sulfur species.
Mg isotopic composition of carbonate: insight from speleothem Geochim. Cosmochim. Acta 43, 869–884.
formation. Earth Planet. Sci. Lett. 201, 105–115. Mottl M. J. and Wheat C. G. (1994) Hydrothermal circulation
Galy A., Yoffe O., Janney P. E., Williams R. W., Cloquet C., Alard through mid-ocean ridge flanks: fluxes of heat and magnesium.
O., Halicz L., Wadhwa M., Hutcheon I. D., Ramon E. and Geochim. Cosmochim. Acta 58, 2225–2237.
Carignan J. (2003) Magnesium isotope heterogeneity of the Murray J. W., Leborgnelt R. and Dandonneau Y. (1997) JGOFS
isotopic standard SRM980 and new reference materials for studies in the equatorial Pacific. Deep Sea Res. Part II Top.
magnesium-isotope-ratio measurements. J. Anal. At. Spectrom. Stud. Oceanogr. 44, 1759–1763.
18, 1352–1356. Nelder J. A. and Mead R. (1965) A simplex method for function
German C. R. and Seyfried, Jr., W. E. (2014) Hydrothermal minimization. Comput. J. 7, 308–313.
processes. In Treatise on Geochemistry (ed. K. K. Turekian). Oelkers E. H., Butcher R., Pogge von Strandmann P. A. E.,
Elsevier, Oxford, England, pp. 191–233. Schuessler J. A., von Blankenburg F., Snæbjörnsdóttir S. Ó.,
Gudbrandsson S., Wolff-Boenisch D., Gı́slason S. R. and Oelkers Mesfin K., Sif Aradóttir E., Gunnarsson I., Sigfússon B.,
E. H. (2011) An experimental study of crystalline basalt Gunnlaugsson E., Matter J. M., Stute M. and Gislason S. R.
dissolution from 26pH611 and temperatures from 5 to 75°C. (2018) Using stable Mg isotope signatures to assess the fate of
Geochim. Cosmochim. Acta 75, 5496–5509. magnesium during the in situ mineralisation of CO2 and H2S at
Guo B., Zhu X., Dong A., Yan B., Shi G. and Zhao Z. (2019) Mg the CarbFix site in SW-Iceland. Geochim. Cosmochim. Acta
isotopic systematic and geochemical applications: a critical 245, 542–555.
review. J. Asian Earth Sci.. Oelkers E. H. and Gı́slason S. R. (2001) The mechanism, rates and
Higgins J. A. and Schrag D. P. (2010) Constraining magnesium consequences of basaltic glass dissolution: I. An experimental
cycling in marine sediments using magnesium isotopes. study of the dissolution rates of basaltic glass as a function of
Geochim. Cosmochim. Acta 74, 5039–5053. aqueous Al, Si and oxalic acid concentration at 25 °C and pH=
Higgins J. A. and Schrag D. P. (2012) Records of Neogene 3 and 11. Geochim. Cosmochim. Acta 65, 3671–3681.
seawater chemistry and diagenesis in deep-sea carbonate Parkhurst D. L. and Appelo C. A. J. (2013) Description of input
sediments and pore fluids. Earth Planet. Sci. Lett. 357, 386–396. and examples for PHREEQC version 3 - A computer program
Higgins J. A. and Schrag D. P. (2015) The Mg isotopic compo- for speciation, batch-reaction, one-dimensional transport, and
sition of Cenozoic seawater – evidence for a link between Mg- inverse geochemical calculations. In U.S. Geological Survey
M. Voigt et al. / Geochimica et Cosmochimica Acta 272 (2020) 21–35 35

techniques and methods, modeling techniques, groundwater U. Steffen M. (1990) A simple method for monotonic interpolation in
S. Geological Survey, Denver, Colorado. one dimension. Astron. Astrophys. 239, 443.
Paton C., Woodhead J. D., Hellstrom J. C., Hergt J. M., Greig A. Teng F.-Z. (2017) Magnesium isotope geochemistry. Rev. Mineral.
and Maas R. (2010) Improved laser ablation U-Pb zircon Geochem. 82, 219–287.
geochronology through robust downhole fractionation correc- Teng F.-Z., Li W.-Y., Ke S., Yang W., Liu S.-A., Sedaghatpour F.,
tion. Geochem. Geophys. Geosystems 11, Q0AA06. Wang S.-J., Huang K.-J., Hu Y. and Ling M.-X., et al. (2015)
Pearce C. R., Saldi G. D., Schott J. and Oelkers E. H. (2012) Magnesium isotopic compositions of international geological
Isotopic fractionation during congruent dissolution, precipita- reference materials. Geostand. Geoanalytical Res. 39, 329–339.
tion and at equilibrium: Evidence from Mg isotopes. Geochim. Teng F.-Z., Li W.-Y., Rudnick R. L. and Gardner L. R. (2010)
Cosmochim. Acta 92, 170–183. Contrasting lithium and magnesium isotope fractionation
Pogge von Strandmann P. A. E., Burton K. W., James R. H., van during continental weathering. Earth Planet. Sci. Lett. 300,
Calsteren P., Gı́slason S. R. and Sigfússon B. (2008) The 63–71.
influence of weathering processes on riverine magnesium Tipper E. T., Galy A. and Bickle M. J. (2006a) Riverine evidence
isotopes in a basaltic terrain. Earth Planet. Sci. Lett. 276, for a fractionated reservoir of Ca and Mg on the continents:
187–197. implications for the oceanic Ca cycle. Earth Planet. Sci. Lett.
Pogge von Strandmann P. A. E., Elliott T., Marschall H. R., Coath 247, 267–279.
C., Lai Y.-J., Jeffcoate A. B. and Ionov D. A. (2011) Variations Tipper E. T., Galy A., Gaillardet J., Bickle M. J., Elderfield H. and
of Li and Mg isotope ratios in bulk chondrites and mantle Carder E. A. (2006b) The magnesium isotope budget of the
xenoliths. Geochim. Cosmochim. Acta 75, 5247–5268. modern ocean: constraints from riverine magnesium isotope
Pogge von Strandmann P. A. E., Opfergelt S., Lai Y.-J., Sigfússon ratios. Earth Planet. Sci. Lett. 250, 241–253.
B., Gı́slason S. R. and Burton K. W. (2012) Lithium, Voigt M., Marieni C., Clark D. E., Gı́slason S. R. and Oelkers E.
magnesium and silicon isotope behaviour accompanying H. (2018a) Evaluation and refinement of thermodynamic
weathering in a basaltic soil and pore water profile in Iceland. databases for mineral carbonation. Energy Procedia 146, 81–91.
Earth Planet. Sci. Lett. 339–340, 11–23. Voigt M., Pearce C. R., Baldermann A. and Oelkers E. H. (2018b)
Pogge von Strandmann P., Forshaw J. and Schmidt D. N. (2014) Stable and radiogenic strontium isotope fractionation during
Modern and Cenozoic records of seawater magnesium from hydrothermal seawater-basalt interaction. Geochim. Cos-
foraminiferal Mg isotopes. Biogeosciences 11, 5155–5168. mochim. Acta 240, 131–151.
Ryu J.-S., Vigier N., Decarreau A., Lee S.-W., Lee K.-S., Song H. White W. M. and Klein E. M. (2014) Composition of the Oceanic
and Petit S. (2016) Experimental investigation of Mg isotope Crust. In Treatise on Geochemistry (ed. K. K. Turekian).
fractionation during mineral dissolution and clay formation. Elsevier, Oxford, pp. 457–496.
Chem. Geol. 445, 135–145. Wimpenny J., Colla C. A., Yin Q.-Z., Rustad J. R. and Casey W.
Schott J., Mavromatis V., Fujii T., Pearce C. R. and Oelkers E. H. H. (2014) Investigating the behaviour of Mg isotopes during the
(2016) The control of carbonate mineral Mg isotope compo- formation of clay minerals. Geochim. Cosmochim. Acta 128,
sition by aqueous speciation: theoretical and experimental 178–194.
modeling. Chem. Geol. 445, 120–134. Wimpenny J., Gı́slason S. R., James R. H., Gannoun A., Pogge
Seewald J. S. and Seyfried, Jr., W. E. (1990) The effect of Von Strandmann P. A. E. and Burton K. W. (2010) The
temperature on metal mobility in subseafloor hydrothermal behaviour of Li and Mg isotopes during primary phase
systems: constraints from basalt alteration experiments. Earth dissolution and secondary mineral formation in basalt.
Planet. Sci. Lett. 101, 388–403. Geochim. Cosmochim. Acta 74, 5259–5279.
Seyfried, Jr., W. E. and Bischoff J. L. (1981) Experimental Young E. D., Galy A. and Nagahara H. (2002) Kinetic and
seawater-basalt interaction at 300°C, 500 bars, chemical equilibrium mass-dependent isotope fractionation laws in
exchange, secondary mineral formation and implications for nature and their geochemical and cosmochemical significance.
the transport of heavy metals. Geochim. Cosmochim. Acta 45, Geochim. Cosmochim. Acta 66, 1095–1104.
135–147. Ziegler K., Chadwick O. A., Brzezinski M. A. and Kelly E. F.
Seyfried, Jr., W. E. and Bischoff J. L. (1977) Hydrothermal (2005) Natural variations of d30Si ratios during progressive
transport of heavy metals by seawater: The role of seawater/ basalt weathering, Hawaiian Islands. Geochim. Cosmochim.
basalt ratio. Earth Planet. Sci. Lett. 34, 71–77. Acta 69, 4597–4610.
Seyfried, Jr., W. E. and Bischoff J. L. (1979) Low temperature
basalt alteration by sea water: an experimental study at 70 °C
Associate editor: Damon A.H. Teagle
and 150 °C. Geochim. Cosmochim. Acta 43, 1937–1947.
Seyfried, Jr., W. E. and Mottl M. J. (1982) Hydrothermal
alteration of basalt by seawater under seawater-dominated
conditions. Geochim. Cosmochim. Acta 46, 985–1002.

You might also like