Accepted Manuscript Diffusion of hydrous species in model basaltic melt Li Zhang, Xuan Guo, Qinxia Wang, Jiale Ding, Huaiwei Ni PII: DOI: Reference:
S0016-7037(17)30427-1 http://dx.doi.org/10.1016/j.gca.2017.07.019 GCA 10381
To appear in:
Geochimica et Cosmochimica Acta
Received Date: Revised Date: Accepted Date:
28 October 2016 6 June 2017 11 July 2017
Please cite this article as: Zhang, L., Guo, X., Wang, Q., Ding, J., Ni, H., Diffusion of hydrous species in model basaltic melt, Geochimica et Cosmochimica Acta (2017), doi: http://dx.doi.org/10.1016/j.gca.2017.07.019
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Diffusion of hydrous species in model basaltic melt
Li Zhang1,2, Xuan Guo1, Qinxia Wang1, Jiale Ding1, Huaiwei Ni1,* 1
CAS Key Laboratory of Crust-Mantle Materials and Environments, School of Earth and Space Sciences, University of Science and Technology of China, Hefei 230026, China 2
Bayerisches Geoinstitut, Universität Bayreuth, 95440 Bayreuth, Germany
Corresponding author: Huaiwei Ni E-mail:
[email protected] Tel. 86-551-63600297
For Geochimica et Cosmochimica Acta on September 28th, 2016
1
Abstract Water diffusion in Fe-free model basaltic melt with up to 2 wt% H2O was investigated at 1658–1846 K and 1 GPa in piston-cylinder apparatus using both hydration and diffusion couple techniques.
Diffusion profiles measured by FTIR are
consistent with a model in which both molecular H2O (H2Om) and hydroxyl (OH) contribute to water diffusion.
OH diffusivity is roughly 13% of H2Om diffusivity,
showing little dependence on temperature or water concentration.
Water diffusion is
dominated by the motion of OH until total H2O (H2Ot) concentration reaches 1 wt%. The dependence of apparent H2Ot diffusivity on H2Ot concentration appears to be overestimated by a previous study on MORB melt, but H2Ot diffusivity at 1 wt% H2Ot in basaltic melt is still greater than those in rhyolitic to andesitic melts.
The
appreciable contribution of OH to water diffusion in basaltic melt can be explained by enhanced mobility of OH, probably associated with the development of free hydroxyl bonded with network-modifying cations, as well as higher OH concentration. Calculation based on the Nernst-Einstein equation demonstrates that OH may serve as an effective charge carrier in hydrous basaltic melt, which could partly account for the previously observed strong influence of water on electrical conductivity of basaltic melt.
2
1. INTRODUCTION
Water is one of the most abundant and pivotal volatile component in terrestrial and lunar magma.
Water diffusivity is a critical parameter for quantitative modeling of
water loss or gain of volcanic glass (Saal et al., 2008; Giachetti and Gonnermann, 2013), bubble growth and resorption (Proussevitch and Sahagian, 1998; McIntosh et al., 2014) and explosive volcanic eruption (Ruprecht and Bachmann, 2010; Lloyd et al., 2014).
Furthermore, water is present in silicate melts in the form of both
molecular H2O (H2Om) and hydroxyl (OH), which are different with respect to both their concentration and mobility (Zhang et al., 1991; Ni et al., 2013).
This
multi-species diffusion problem is theoretically inspiring in terms of the coupling of diffusion and speciation reaction as well as the relationship between the apparent total water (H2Ot) diffusivity and the diffusivities of the two hydrous species. Most of previous studies on water diffusion in silicate melts focused on felsic to intermediate melt compositions (e.g., Shaw, 1974; Zhang et al., 1991; Nowak and Behrens, 1997; Zhang and Behrens, 2000; Liu et al., 2004a,b; Behrens et al., 2004; Ni and Zhang, 2008; Ni et al., 2009a,b; Schmidt et al., 2013; Fanara et al., 2013; Ni et al., 2013; Persikov et al., 2014).
Zhang et al. (1991) investigated dehydration in
rhyolitic melt within a limited H2Ot range (< 2 wt%) and concluded that H2Om is the mobile species while OH diffusivity is negligible.
Zhang and Behrens (2000)
showed that if H2Om diffusivity was assumed to increase exponentially with H2Ot
3
concentration, the basic idea in Zhang et al. (1991) could be applicable for higher H2Ot concentration.
This speciation-based diffusion model found great success in
later studies for rhyolitic and dacitic melts (Liu et al., 2004a; Ni and Zhang, 2008; Ni et al., 2009b).
However, in andesitic melt, Behrens et al. (2004) found that some
H2Ot diffusion profiles were better fit by constant H2Ot diffusivity than by H2Ot-dependent diffusivity, which implies a role for OH diffusion.
Ni et al. (2013)
resolved the contribution of OH to water diffusion in andesitic melt and found OH diffusivity to be 10% – 20% of H2Om diffusivity at low H2Ot concentration.
These
results suggest that OH diffusion is probably more significant in depolymerized melts than in polymerized melts. With respect to the mafic systems, Zhang and Stolper (1991) investigated water diffusion in a mid-ocean ridge basalt (MORB) melt with low H2Ot content (<0.4 wt%) at 1573–1773 K and 1 GPa.
Their results were roughly consistent with the model of
H2Ot diffusivity being proportional to H2Ot concentration, which was attributed to an H2Om-dominated diffusion mechanism similar to what operates in rhyolitic melt (Zhang et al., 1991).
This conclusion was echoed by Okumura and Nakashima
(2006) in their investigation of water diffusion in a basaltic glass at ambient pressure. However, Zhang and Ni (2010) found that the diffusivity data in Zhang and Stolper (1991) extrapolated to 1 wt% H2Ot are significantly off the compositional trend constrained by experimental data for felsic and intermediate melts.
Furthermore, in
view of the findings in Behrens et al. (2004) and Ni et al. (2013) regarding OH
4
diffusion in andesitic melt, the role of OH diffusion might be even more significant in the more depolymerized basaltic melt, therefore challenging the H2Om-dominated mechanism proposed by Zhang and Stolper (1991).
The only other study on water
diffusion in basaltic melt is from Persikov et al. (2010), but their “haplobasalt” composition (Albite45-Diopside37-Wollastonite18) is clearly not a satisfactory analogue of naturally occurred basalts. To collect more reliable data of water diffusivity and evaluate the importance of OH diffusion in basaltic melt, we have carried out hydration and diffusion couple experiments at 1658–1846 K and 1 GPa for basaltic melt with up to 2 wt% H2Ot. The experimental results are useful not only for more robustly constraining the dependences of water diffusivity on H2Ot concentration and melt composition, but also for clarifying the underlying physical mechanism that governs water diffusion in silicate melts.
Their implication for electrical conduction in hydrous basaltic melts
will also be discussed.
2. EXPERIMENTAL AND ANALYTICAL METHODS 2.1. Starting material For synthesis of starting glasses used in diffusion experiments, the targeting composition was taken from the average composition of basalt in the calc-alkaline series of island-arc volcanic suites (Hess, 1989).
But to avoid iron-associated
complexities, including Fe-Pt alloying, Fe valence change and reduction of glass
5
transparency, we opted for a haplobasaltic composition by replacing Fe2+ with Ca and Mg while keeping the original molar Ca/Mg ratio.
Reagent-grade oxides and
carbonates were mixed and fused twice at 1823 K in a 1-bar furnace to obtain the model anhydrous basaltic glass.
The glass composition as analyzed by electron
microprobe (Table 1) is similar to the MORB used in Zhang and Stolper (1991) but differs significantly from the haplobasalt studied in Persikov et al. (2010). A series of hydrous glasses were prepared by fusing the mixture of anhydrous glass powder and distilled water (sealed in Pt capsules) at 1623 K and 0.5 GPa for 10 h in two end-load piston cylinder apparatuses, one located at the University of Science and Technology of China (USTC) and the other at the Bayerisches Geoinstitut (BGI), Germany.
The highest permissible water content without causing crystallization
during quench was found to be about 2 wt%. All of the synthesized glasses are transparent, with the hydrous glasses being crystal- and bubble-free while the anhydrous glass containing a small amount of air bubbles.
As determined from FTIR analysis, the H2Ot concentration of the
nominally anhydrous glass is 0.03 wt%, and the hydrous glasses contain 0.1–2.0 wt% homogeneously distributed water.
2.2. Diffusion experiments Glass cylinders of 2.6 mm diameter were prepared for diffusion experiments.
In
one set of experiments, aimed at determining OH diffusivity at extremely low H2Ot
6
concentration, a piece of 4-mm long anhydrous glass cylinder with both ends polished was sealed in a Pt capsule (3 mm diameter and 0.15 mm wall thickness).
The other
set of experiments following the double diffusion couple approach developed by Ni et al. (2013) was designed for water diffusivity over a broad range of H2Ot concentration. For each diffusion couple sealed in a Pt capsule, a 2.5 mm-long hydrous glass cylinder was placed on top of an anhydrous or less hydrous glass cylinder of 1.5 mm length, with their polished surface touching each other.
Two diffusion couples with
different combination of water contents investigated in a single experimental run allow to more accurately constrain the dependence of water diffusivity on H 2Ot concentration (Ni et al., 2013). Diffusion experiments were carried out in 3/4˝ end-load piston cylinder apparatuses at the USTC and at the BGI.
The sample assembly was composed of
talc, pyrex glass, graphite heater and crushable alumina or MgO.
A type-S
thermocouple (Pt-Pt95Rh5) atop the Pt capsule served to monitor the temperature, and the temperature at the sample was estimated to be 15 K higher than the thermocouple reading.
A “piston-out” procedure was applied for pressurization, with a friction
correction of 18% on the nominal pressure.
The sample charge was first relaxed at
573 K and 1 GPa for at least 3 h, and then rapidly (in ~60 s) heated to a target temperature in the range of 1593–1831 K.
After a dwell time of 2–4 min to allow
for diffusive exchange of water, power was immediately turned off to achieve rapid quench with an onset cooling rate of ~80 K/s.
7
With computer-recorded thermal
history, the diffusion that occurred during heating and cooling was accounted for by calculating an effective experimental duration, as described in Zhang and Behrens (2000), with an estimated activation energy of 220 kJ/mol for water diffusion. After the diffusion experiment, the recovered sample was mounted into epoxy resin and doubly polished to 150–250 m thickness for FTIR analyses.
The samples
(especially those from the hydration experiments) often contain some cracks perpendicular to the cylindrical axis (Figs. S1 and S2), causing some complexity in the measurement of diffusion profiles and water diffusivity.
2.3. FTIR analyses Initial water contents of the starting glasses were determined by a PerkinElmer Frontier FTIR spectrometer at the USTC.
The MIR setup (MIR source + KBr
beamsplitter + MIRTGS) was adopted for a majority of samples.
For the O-H
stretching band at 3550 cm–1, its peak height was used to calculate water content according to the Lambert-Beer law, with a molar absorptivity of 63 L·mol–1·cm–1 (Dixon et al., 1995).
For glasses with >1.4 wt% H2Ot, this strong MIR band easily
got oversaturated, therefore the two combination bands at 4520 cm–1 and at 5200 cm–1 were measured using the NIR setup (NIR source + CaF2 beamsplitter + NIRTGS) to obtain OH concentration and H2Om concentration, respectively.
Calibration of the
NIR bands was also from Dixon et al. (1995), but the procedure of baseline correction was simplified by applying two linear tangential baselines.
8
This simplification may
lead to <10% uncertainty in the calculation of H2Ot concentration, but will not cause a major effect on the determination of water diffusivity, which is mostly constrained by the length of diffusion profile. Diffusion profiles preserved in the quenched samples were measured along the centerline by a Spotlight 200 microscope system (involving an MCT detector) attached to the PerkinElmer Frontier FTIR spectrometer. consists of 70 to 120 points.
Each profile typically
At each point, an FTIR spectrum was collected using a
rectangular aperture of 10 m width and 200 m length, with 64 scans through 6000–2000 cm–1.
The obtained infrared spectra were processed similarly as done in
obtaining the initial water contents of glasses.
Our diffusion profiles were at least
800 m long, implying a minimal convolution effect in FTIR microspectroscopy.
3. EXPERIMENTAL RESULTS We conducted seven diffusion couple experiments and four hydration experiments at 1608–1846 K (Table 2).
The two samples from Run# HBS-BGI-DC3 at 1608 K
and one from Run# HBS-BGI-DC4 (DC4b) at 1658 K were severely crystallized, forbidding diffusion profile acquisition.
The lower bound of allowable experimental
temperature was therefore limited at 1658 K.
For most other samples, the water
contents at the flat regions of diffusion profiles remain the same as the initial water contents (Figs. S3 and S4), confirming that our data can be treated as a diffusion problem in infinite (for diffusion couple) or semi-infinite (for hydration) medium.
9
In
one sample from Run# HBS-USTC-DC7 (DC7a) at 1846 K, the highest temperature explored, the anhydrous half was undesirably overhydrated to >0.14 wt% H2Ot. After dismissing the problematic samples, in total 14 diffusion profiles were successfully obtained, 10 from the diffusion couple experiments and 4 from the hydration experiments (Figs. S1-S4).
Mole fraction of total water (X) is related to
H2Ot concentration (C in wt%) by X = C/18.015/[C/18.015+(100-C)/35.03], in which 35.03 g/mol is the molar mass of anhydrous basalt on a single oxygen basis. At the two ends of the profiles from diffusion couple experiments (i.e., outside the flat regions in Fig. S3), we observed modest degree of dehydration of the hydrous half and hydration of the anhydrous half over a short distance, as already documented in previous studies (e.g., Zhang and Behrens, 2000).
Alteration in H2Ot concentration
also occurred near the sidewalls of Pt capsules.
Nevertheless, the initial water
contents were exactly retained at the flat regions of diffusion profiles measured along the centerline.
Therefore, we conclude that the observed changes in water content at
the rims of samples did not affect the diffusive exchange between the two halves. Diffusion profiles at the two ends of samples from the 4 hydration experiments were found to be always symmetric with respect to center of sample.
The H2Ot
concentration at the melt-capsule interface was elevated consistently to ~600 ppm over the investigated temperature range.
In view of the high intrinsic water fugacity
in the sample assembly caused by the decomposition of talc, we suggest that the gained water originated from the ingassing of hydrogen through the Pt capsule in the
10
form of H2, in addition to any trapped moisture in the capsule.
The incoming H2 was
expected to be quickly oxidized to H2O at the melt-capsule interface, presumably by trapped O2 inside the capsule and trace of ferric iron in the melt.
4. DISCUSSION AND APPLICATION 4.1. Diffusivity model and diffusion mechanism We first tried to fit the measured diffusion profiles with error function, which corresponds to a model of water diffusivity independent of H2Ot concentration. While acceptable fitting quality is obtained for the hydration profiles (Fig. S4), those profiles from diffusion couple experiments involving high H2Ot concentration cannot be well fitted, as shown in Fig. 1a.
Furthermore, for several experiments at similar
temperature, the diffusivity extracted from error function fits shows a positive correlation with average H2Ot concentration.
These two findings suggest that water
diffusivity in silicate melt must vary with H2Ot concentration.
Persikov et al. (2010)
assumed H2Ot diffusivity to be an exponential function of H2Ot concentration. However, this simple model also produces considerable misfit (Fig. 1a). Zhang et al. (1991) developed a mechanistic water diffusivity model for rhyolitic melt.
There are three assumptions in their model: (1) molecular H2O (H2Om) is the
dominating diffusing species; (2) the contribution of OH to water diffusion is negligible; (3) local equilibrium is always maintained for the H2O speciation reaction, H2Om + O = 2OH (here O denotes an anhydrous oxygen atom).
11
Zhang and Behrens
(2000) further showed H2Om diffusivity to be an exponential function of H2Ot concentration rather than being constant.
However, for andesitic melt, Ni et al.
(2013) found significant misfit at low H2Ot content by using this speciation-based model, and pointed out this misfit was due to neglect of OH diffusion.
They
provided a solution by developing a modified speciation-based H2O diffusion model, which assumes (a) H2Om diffusivity is an exponential function of H2Ot concentration; and (b) OH diffusivity is a non-zero constant. We found that our diffusion couple profiles could also be fit well by this model.
Moreover, the dependence of H2Om
diffusivity on H2Ot concentration appears to be weak and almost negligible. Therefore, we set both H2Om diffusivity and OH diffusivity to be constant at given temperature and pressure.
Diffusion couple profile can be fitted excellently with this
modified speciation-based model (Fig. 1a).
Furthermore, the dependence of H2Ot
diffusivity on H2Ot concentration yielded by this model agrees well with the result from Boltzmann-Matano analysis (Fig. 1b), which makes no a priori assumption on the diffusivity-concentration relationship. For the modified-speciation based diffusion model, its mathematical formulation is as follows (Zhang, 1999; Wang et al., 2009),
¶X ¶t
¶ æ ¶X ö DH O ç ¶x è 2 t ¶x ÷ø
K
¶ XH O ¶ æ 2 m ç DH 2Om ¶x è ¶x
DOH
2 X OH 2453 exp 1.547 X H 2 Om X O T
12
¶ X OH ö ÷ 2¶x ø
dX DH2Ot DH2Om 1 OH 2dX
dX OH DOH 2dX
dX OH 1 2X 2dX 4 X ( X 1)(1 4 / K ) 1
The definitions of symbols are summarized in Table 3.
Essentially, equation (1) is
the governing diffusion equation, equation (2) gives the temperature-dependent equilibrium constant of the speciation reaction H2Om + O = 2OH, and equations (3) and (4) relate H2Ot diffusivity to the diffusivities of the two hydrous species. Because the equilibrium water speciation in basaltic melt is currently still unknown, the expression of K in equation (2) is borrowed from the result for andesitic melt (Ni et al., 2009a).
All the profiles from diffusion couple experiments are fitted well
using this model (Fig. S3), and the obtained diffusivities of the hydrous species alongside the DOH/DH2Om ratio are reported in Table 4. It should be noted that for the hydration profiles, because H2Ot concentration is extremely low (X < 0.0012), the differential term in equation (4) essentially equals unity and the extracted DH2Ot diffusivity approximates OH diffusivity, which is also reported in Table 4.
Both DH2Om and DOH data line up nicely in the Arrhenius plot
with similar slope (Fig. 2), except that Run# HBS-BGI-DC5 at 1828 K gives anomalously lower diffusivity than most data.
The reason for this outlier run is not
clear, but may be associated with the longer time (~2 min) taken for heating to the
13
target temperature (and therefore leading to the large correction on experimental duration, see Table 2).
Nevertheless, the OH diffusivities obtained from the two sets
of experiments, hydration and diffusion couple, generally show excellent agreement. After dismissing the outliers from Run# HBS-BGI-DC5, least squares fitting yields the following expressions for the diffusivities of hydrous species in basaltic melt,
and
lnDOH = (–6.731 ± 0.796) + (–26095 ± 1387)/T,
(5)
lnDH2Om = (–4.539 ± 1.047) + (–26304 ± 1819)/T,
(6)
where DOH and DH2Om are both in m2/s.
The fitting results yield an activation energy
of ~220 kJ/mol for both OH diffusion and H2Om diffusion.
The DOH/DH2Om ratio is
nearly constant (~0.13) within the investigated temperature range. One would expect that the diffusion of OH (a charged species) should involve a higher activation energy than for H2Om (a neutral species).
However, different
transport processes are increasingly coupled with increasing temperature and pressure, and will eventually converge to a uniform rate and mechanism (Ni et al., 2015). In vibrant hydrous basaltic melt at high temperature, the interaction between H2Om and local melt structure could be stronger than conventionally thought.
The energy
barrier for the transport of H2Om to overcome may not be significantly lower than that for OH.
Furthermore, due to the limited temperature range in our experiments, the
precision of inferred activation energies may not allow to resolve a small difference between H2Om and OH.
14
4.2. H2Ot diffusivity compared with previous studies H2Ot diffusivity in basaltic melt can now be calculated by combining equations (5)–(6) with equations (2)–(4), with an example shown in Fig. 3 for 1846 K and 1 GPa.
Clearly, in the low H2Ot concentration range (<0.1 wt%) in basaltic melt,
water diffusion is overwhelmingly dominated by OH, OH diffusivity and H2Ot diffusivity are essentially identical, which corroborates our practice of using error function to fit the hydration profiles.
The contribution of OH continues to exceed
that of H2Om until H2Ot concentration increases to 1 wt%.
The conventional wisdom
of H2Om domination becomes valid only at even higher H2Ot concentration. For the convenience of application to magmatic processes, we provide the following equation for approximate calculation of H2Ot diffusivity (DH2Ot is in m2/s) as a function of temperature (T in K) and water content (C in wt%):
DH2Ot e
6.731
26095 T
1.36 0.126 1 Ce
C+
2274 182.7 C 290.6 C T
This empirical expression is applicable for temperature in the range of 1658–1846 K and water content up to 2 wt%, and is able to reproduce the calculation results using equations (2)–(6) within 1% relative. Compared with existing experimental data, at least for T > 1300 K, H2Ot diffusivity at 1 wt% H2Ot in silicate melts follows the sequence rhyolite < dacite < andesite < basalt (Fig. 4).
Furthermore, there appears to be a consistent trend for
activation energy, increasing from 110 kJ/mol for rhyolite, to 144 kJ/mol for dacite, to 168 kJ/mol for andesite, and to 189 kJ/mol for basalt. 15
Zhang and Ni (2010)
developed a general expression of H2Ot diffusivity at 1 wt% H2Ot for rhyolitic to andesitic melts, but they found that the data for MORB melt in Zhang and Stolper (1991) could not be described by that equation.
By contrast, our data significantly
fall below those in Zhang and Stolper (1991) but are in good agreement with the model prediction of Zhang and Ni (2010), as shown in Fig. 4a.
We suspect that
Zhang and Stolper (1991) probably overestimated H2Ot diffusivity at 1 wt% H2Ot in basaltic melt, which was extrapolated from their experimental data at 0.2 wt% H2Ot assuming a proportional relationship.
Indeed, when the two data sets are compared
at 0.2 wt% H2Ot, their agreement appears to be significantly improved (Fig. 4b). The good agreement between our results and previous studies also corroborates the validity of using an Fe-free analogue for basaltic melt.
4.3. Perspective from melt structure The contribution of OH to water diffusion was not found by earlier studies on rhyolitic and dacitic melts (e.g., Zhang and Behrens, 2000), but was only resolved recently for more depolymerized melts, including andesitic melt in Ni et al. (2013) and basaltic melt in the present study.
There are at least three considerations from
the perspective of melt structure that may help to explain this difference. Firstly, andesitic and basaltic melts, with a lower degree of polymerization, have lower viscosity than rhyolitic and dacitic melts (Hess and Dingwell, 1996; Richet et al., 1996; Giordano and Dingwell, 2003a; Whittington et al., 2009a; Ni et al., 2015).
16
Diffusivities of atomic species, especially those with low mobility, are strongly coupled with viscosity.
For example, the Eyring equation declares that Si diffusivity
is inversely proportional to melt viscosity (Glasstone et al., 1941; Zhang et al., 2010), and experimental data show that Si diffusivity in basaltic melt is higher than that in rhyolitic melt by nearly 4 orders of magnitude (Baker, 1992; Lesher et al., 1996). OH diffusivity is also expected to increase significantly from rhyolitic to basaltic melt. Furthermore, the equilibrium constant of the water speciation reaction appears to increase from rhyolitic to andesitic melt (Zhang and Ni, 2010).
This trend may
continue to hold toward basaltic melt although it awaits experimental confirmation. This means that for the same temperature, pressure and H2Ot concentration, there is more OH present in the more depolymerized melts.
The differential term in equation
(4), which corresponds to the weight for the relative contribution of OH to water diffusion, also increases with increasing K. Last but perhaps more importantly, NMR spectroscopic studies have demonstrated that in depolymerized melts, OH bonds not only with network-forming cations such as Si and Al but also with network-modifier cations such as Ca and Mg (Xue and Kanzaki, 2004, 2008; Xue, 2009). The latter kind of OH, often referred to as free hydroxyl or OH– (Xue and Kanzaki, 2004), also becomes more abundant with increasing temperature (Moretti et al., 2014).
Due to its weaker bonding with cation,
free hydroxyl is expected to have higher mobility than Si-OH or Al-OH.
17
All the three microscopic mechanisms above can rationalize why OH plays a more perceptible role in water diffusion in the more depolymerized melts.
In essence, OH
in depolymerized melts is more mobile and more abundant than that in polymerized melts.
4.4. OH contribution to electrical conduction Silicate melts are ionic conductors, and electrical conductivity of silicate melts is useful for interpreting magnetotelluric data and probing the status of molten regions (such as water content of magma) in Earth’s interior (e.g., Gaillard, 2004; Ni et al., 2011).
Because OH is a charged species, it may serve as a potential charge carrier in
hydrous basaltic and andesitic melts.
The contribution of an ionic species i to
electrical conduction can be evaluated from diffusivity through the Nernst-Einstein equation
i =
1 Di ci ( zi F )2 Hi RT
where i is the contribution of the ionic species to electrical conductivity (in S/m), Di, ci and zi are the diffusivity (in m2/s), the concentration (in mol/m3), and the valence of the species, and Hi is the Haven ratio depending on the diffusion mechanism. For comparison, we use Na, the ion with typically the highest diffusivity, as a reference.
At low H2Ot concentration, Na diffusivity is similar in basaltic and
andesitic melts (Lowry et al., 1982), and is higher than OH diffusivity in andesitic melt by more than an order of magnitude (Fig. 5). But for basaltic melt, Na diffuses 18
faster than OH diffusivity by only a factor of 2–4.
In hydrous melt, the mobility of
ions will be enhanced in correspondence to reduced viscosity.
Watson (1981)
showed that Na diffusivity in rhyolitic melt is nearly doubled upon the addition of 3.5 wt% water. The influence of water on Na diffusivity in basaltic melt is probably similar or even weaker.
OH diffusivity may also increase with increasing H2Ot
concentration although this effect cannot be easily resolved.
Overall, it is reasonable
to expect that the DOH/DNa ratio is roughly independent of H2Ot concentration.
If
one assumes that OH and Na diffuse by a similar mechanism so that their Haven ratio is cancelled out, the contribution of OH to electrical conduction relative to Na can be expressed as follows:
OH DOH cOH Na DNa cNa
The concentration of OH at different H2Ot can be calculated from the equilibrium constant K.
As shown in Fig. 6, OH contribution is generally no more than several
percent of Na contribution in andesitic melt unless for very high H2Ot concentration. But in basaltic melt, OH contribution can become even greater than Na contribution, which makes OH an effective charge carrier in hydrous basaltic melt. It has been noticed that water enhances electrical conductivity of basaltic and andesitic melts more significantly than the effect for rhyolitic and dacitic melts (Ni et al., 2011; Guo et al., 2016; Guo et al., 2017). An explanation given in Guo et al. (2016) suggested that in the depolymerized melts network-modifying cations other than Na (e.g., Mg and Ca) play a major role in electrical conduction, and that adding 19
water mobilizes these cations more effectively than it does for Na, the predominant charge carrier in rhyolitic melt.
Our analysis above provides a second mechanism.
That is, at least for basaltic melt, adding water not only enhances the mobility of cations, OH itself (an anion) also contributes to carrying electrical charge.
4.5. Application to bubble growth in magma In addition to melt viscosity and water solubility, water diffusivity is another critical parameter in determining the rate of bubble growth, an important process in explosive volcanic eruption (Proussevitch and Sahagian, 1998; Liu and Zhang, 2000; Lensky et al., 2004; Houghton and Gonnermann, 2008).
Based on the bubble
growth model developed by Proussevitch and Sahagian (1998), slightly modified by Liu and Zhang (2000), we compare bubble growth in basaltic melt and rhyolitic melt with 2 wt% H2O and 1573 K and 50 MPa, corresponding to a P-T condition in a volcanic conduit (Fig. 7).
The parabolic shape of the growth curves indicates that
bubble growth is controlled by water diffusion instead of by viscous flow.
Higher
water diffusivity in basaltic melt leads to appreciably faster bubble growth than in rhyolitic melt.
This may facilitate separation of the bubbles from basaltic magma
and have a major impact on the eruption dynamics.
5. CONCLUSIONS Water diffusion in silicate melts is conventionally thought to be dominated by
20
neutral H2O molecules rather than the charged OH species.
However, the present
study shows that OH is mobile enough for contributing significantly to water diffusion in basaltic melt, especially at low H2Ot concentration.
In view of the
results from NMR investigations, the high mobility of OH is probably due to the notion that a significant fraction of OH is free hydroxyl bonded with network-modifying cations instead of with Si or Al.
The strong influence of water
on electrical conductivity of basaltic melt can be accounted for, at least partly, by OH serving as an effective charge carrier in addition to cations.
Acknowledgements We thank Youxue Zhang for instructive suggestions and Yan Liang for insightful and constructive comments.
This work was supported by the National Natural
Science Foundation of China (41473058), the Strategic Priority Research Program (B) of the Chinese Academy of Sciences (XDB18000000), and the Fundamental Research Funds for the Central Universities of China.
References Baker D. R. (1992) Tracer diffusion of network formers and multicomponent diffusion in dacitic and rhyolitic melts. Geochim. Cosmochim. Acta 56, 617–631. Behrens H., Zhang Y. and Xu Z. (2004) H2O diffusion in dacitic and andesitic melts. Geochim. Cosmochim. Acta 68, 5139–5150. 21
Bezos A. and Humler E. (2005) The Fe3+/Fe ratios of MORB glasses and their implications for mantle melting. Geochim. Cosmochim. Acta 69, 711-725. Dixon J. E., Stolper E. M. and Holloway J. R. (1995) An experimental study of water and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part I: calibration and solubility models. J. Petrol. 36, 1607–1631. Fanara S., Behrens H. and Zhang Y. (2013) Water diffusion in potassium-rich phonolitic and trachyte melts. Chem. Geol. 346, 149–161. Gaillard F. (2004) Laboratory measurements of electrical conductivity of hydrous and dry silicic melts under pressure. Earth Planet. Sci. Lett. 218, 215–228. Giachetti T. and Gonnermann H. M. (2013) Water in volcanic pyroclast: rehydration or incomplete degassing? Earth Planet. Sci. Lett. 369–370, 317–332. Giordano D. and Dingwell D. B. (2003a) Viscosity of hydrous Etna basalt: Implication for Plinian-style basaltic eruption. Bull. Volcanol. 65, 8–14. Glasstone S., Laidler K. J. and Eyring H. (1941) The Theory of Rate Processes. McGraw-Hill, New York. Guo X., Zhang L., Behrens H. and Ni H. (2016) Probing the status of felsic magma reservoirs: Constraints from the P–T–H2O dependences of electrical conductivity of rhyolitic melt. Earth Planet. Sci. Lett. 433, 54–62. Guo X., Li B., Ni H. and Mao Z. (2017) Electrical conductivity of hydrous andesitic melts pertinent to subduction zones. J. Geophys. Res. 122, 1777–1788. Hess P. C. (1989) Origins of Igneous Rocks. Harvard University Press.
22
Hess K. U. and Dingwell D. B. (1996) Viscosities of hydrous leucogranitic melts: A non-Arrhenius model. Am. Mineral. 81, 1297–1300. Houghton B. F. and Gonnermann H. M. (2008) Basaltic explosive volcanism: constrains from deposits and models. Chem. Erde-Geochem. 68, 117–140. Hui H. and Zhang Y. (2007) Towards a general viscosity equation for natural anhydrous and hydrous silicate melts. Geochim. Cosmochim. Acta 71, 403–416. Lensky N. G., Navon O. and Lyakhovsky V. (2004) Bubble growth during decompression of magma: experimental and theoretical investigation. J. Volcanol. Geotherm. Res. 129, 7–22. Lesher C. E., Hervig R. L. and Tinker D. (1996) Self diffusion of network formers (silicon and oxygen) in naturally occurring basaltic liquid. Geochim. Cosmochim. Acta 60, 405–413. Liu Y. and Zhang Y. (2000) Bubble growth in rhyolitic melt. Earth Planet. Sci. Lett. 181, 251–264. Liu Y., Behrens H. and Zhang Y. (2004b) The speciation of dissolved H 2O in dacitic melt. Am. Mineral. 89, 277–284. Liu Y., Zhang Y. and Behrens H. (2004a) H2O diffusion in dacitic melts. Chem. Geol. 209, 327–340. Lloyd A. S., Ruprecht P., Hauri E. H., Rose W., Gonnermann H. M. and Plank T. (2014) NanoSIMS results from olivine-hosted melt embayments: Magma ascent rate during explosive basaltic eruptions. J. Volcanol. Geotherm. Res. 283, 1–18.
23
Lowry R. K., Henderson P. and Nolan J. (1982) Tracer diffusion of some alkali, alkaline-earth and transition element ions in a basaltic and an andesitic melt, and the implications concerning melt structure. Contrib. Mineral. Petrol. 80, 254–261. McIntosh I. M., Liewellin E. W., Humphreys M. C. S., Nichols A. R. L., Burgisser A., Schipper C. I. and Larsen J. F. (2014) Distribution of dissolved water in magmatic glass records growth and resorption of bubbles. Earth Planet. Sci. Lett. 401, 1–11. Moretti R., Le Losq C. and Neuville D. R. (2014) The amphoteric behavior of water in silicate melts from the point of view of their ionic-polymeric constitution. Chem. Geol. 367, 23–33. Ni H. and Zhang Y. (2008) H2O diffusion models in rhyolitic melt with new high pressure data. Chem. Geol. 250, 68–78. Ni H., Behrens H. and Zhang Y. (2009b) Water diffusion in dacitic melt. Geochim. Cosmochim. Acta 73, 3642–3655. Ni H., Hui H. and Neumann G. S. (2015) Transport properties of silicate melts. Rev. Geophys. 53, 715–744. Ni H., Keppler H. and Behrens H. (2011) Electrical conductivity of hydrous basaltic melts: implications for partial melting in the upper mantle. Contrib. Mineral. Petrol. 162, 637–650. Ni H., Liu Y., Wang L. and Zhang Y. (2009a) Water speciation and diffusion in
24
haploandesitic melts at 743–873 K and 100 MPa. Geochim. Cosmochim. Acta 73, 3630–3641. Ni H., Xu Z. and Zhang Y. (2013) Hydroxyl and molecular H 2O diffusivity in a haploandesitic melt. Geochim. Cosmochim. Acta 103, 36–48. Nowak M. and Behrens H. (1997) An experimental investigation on diffusion of water in haplogranitic melts. Contrib. Mineral. Petrol. 126, 365–376. Okumura S. and Nakashima S. (2006) Water diffusion in basaltic to dacitic glasses. Chem. Geol. 227, 70–82. Persikov E. S., Bukhtiyarov P. G., Nekrasov A. N. and Bondarenko G. V. (2014) Concentration dependence of water diffusion in obsidian and dacitic melts at high pressures. Geochem. Inter. 52, 365–371. Persikov E. S., Newman S., Bukhtiyarov P. G., Nekrasov A. N. and Stolper E. M. (2010) Experimental study of water diffusion in haplobasaltic and haploandesitic melts. Chem. Geol. 276, 241–256. Proussevitch A. A. and Sahagian D. L. (1998) Dynamics and energetics of bubble growth in magmas: analytical formulation and numerical modeling. J. Geophys. Res. 103, 18223–18251. Richet P., Lejeune A.-M., Holtz F. and Roux J. (1996) Water and the viscosity of andesite melts. Chem. Geol. 128, 185–197. Ruprecht P. and Bachmann O. (2010) Pre-eruptive reheating during magma mixing at Quizapu volcano and the implications for the explosiveness of silicic arc
25
volcanoes. Geology 38, 919–922. Saal A. E., Hauri E. H., Cascio M., Van Orman J. A., Rutherford M. C. and Cooper R. F. (2008) Volatile content of lunar volcanic glasses and the presence of water in the Moon’s interior. Nature 454, 192–195. Schmidt B. C., Blum-Oeste N. and Flagmeier J. (2013) Water diffusion in phonolite melts. Geochim. Cosmochim. Acta 107, 220–230. Shaw H. R. (1974) Diffusion of H2O in granite liquids: I. Experimental data; II. Mass transfer in magma chamber. In Geochemical Transport and Kinetics (eds. A. W. Hofmann, B. J. Giletti, H. S. Yoder, and R. A. Yund). Carnegie Inst. Washington Publ., Washington, DC. pp. 139–170. Wang H., Xu Z., Behrens H. and Zhang Y. (2009) Water diffusion in Mount Changbai peralkaline rhyolitic melt. Contrib. Mineral. Petrol. 158, 471–484. Waston E. B. (1981) Diffusion in magmas at depth in the Earth: the effect of pressure and dissolved H2O. Earth Planet. Sci. Lett. 52, 291–301. Whittington A. G., Hellwig B. M., Behrens H., Joachim B., Stechern A. and Vetere F. (2009a) The viscosity of hydrous dacitic liquids: Implication for the rheology of evolving silicic magmas. Bull. Volcanol. 71, 185–199. Xue X. (2009) Water speciation in hydrous silicate and aluminosilicate glasses: direct evidence from
29
Si–1H and
27
Al–1H double-resonance NMR. Am. Mineral. 94,
395–398. Xue X. and Kanzaki M. (2004) Dissolution mechanism of water in depolymerized
26
silicate melts: constrains from 1H and
29
Si NMR spectroscopy and ab initio
calculations. Geochim. Cosmochim. Acta 68, 5027–5057. Xue X. and Kanzaki M. (2008) Structure of hydrous aluminosilicate glasses along the diopside–anorthite join: a comprehensive one- and two-dimensional 1H and
27
Al
NMR study. Geochim. Cosmochim. Acta 72, 2331–2348. Zhang Y. (1999) H2O in rhyolitic glasses and melts: measurement, speciation solubility, and diffusion. Rev. Geophys. 37, 493–516. Zhang Y. and Behrens H. (2000) H2O diffusion in rhyolitic melts and glasses. Chem. Geol. 169, 243–262. Zhang Y. and Ni H. (2010) Diffusion of H, C, and O components in silicate melts. Rev. Mineral. Geochem. 72, 171–225. Zhang Y., Ni H. and Chen Y. (2010) Diffusion data in silicate melts. Rev. Mineral. Geochem. 72, 311–408. Zhang Y. and Stolper E. M. (1991) Water diffusion in a basaltic melt. Nature 351, 306–309. Zhang Y., Stolper E. M. and Wasserburg G. J. (1991) Diffusion of water in rhyolitic glasses. Geochim. Cosmochim. Acta 55, 441–456. Zhang Y., Xu Z., Zhu M. and Wang H. (2007) Silicate melt properties and volcanic eruptions. Rev. Geophys. 45, RG4004. doi: 10.1029/2006RG000216.
27
Table 1 Chemical composition of basaltic glasses (in wt%) used in H2O diffusion studies HBS-dry
MORB
Haplobasalt
(this study)
(Zhang and Stolper, 1991)
(Persikov et al., 2010)
SiO2
51.97
50.60
62.20
TiO2
1.06
1.88
Al2O3
16.32
13.90
FeOt
0.03
12.50
MgO
11.21
6.56
6.79
CaO
15.34
11.40
14.15
Na2O
2.79
2.64
6.34
K2O
0.89
0.17
MnO
0.23
P2O5
0.21
Total H2O (IR) NBO/T
10.50
99.61
100.09
99.98
0.03
0.4
<0.1
0.798
0.826
0.677
Major element composition of HBS-dry was analyzed by a Shimadzu 1600 electron microprobe with a defoucus beam of 15 kV, 10 nA, and 5 m across.
Water content was measured by FTIR
spectroscopy using the calibration for the 3550 cm–1 band by Dixon et al. (1995). NBO/T is the number of non-bridging oxygen atoms per tetrahedrally coordinated cation.
For MORB, Fe3+
accounts for 13.6% of total iron (Bezos and Humler, 2005) and is counted as network-forming cation.
28
Table 2 Experimental conditions of water diffusion experiments at 1 GPa Run #a
Tdwell (K)
Tcorrected (K)
tdwell (s)
teffective b (s)
Initial H2Ot c (wt%)
Final H2Ot d (wt%)
1693
1708±20
132
150
0.33/0.74
0.33/0.74
0.31/0.71
0.31/0.71
0.31/1.77
0.32/1.78
0.28/1.13
0.28/1.11
0.21/1.92
n.a.
Crystallization
0.21/1.04
n.a.
Crystallization
0.16/1.90
0.16/1.92
Partial crystal.
0.16/1.04
n.a.
Crystallization
0.39/1.53
0.39/1.51
0.36/0.95
0.36/0.95
0.03/1.98
0.04/1.98
0.03/1.02
0.03/1.02
0.03/1.98
0.14/2.02
0.03/0.95
0.04/0.95
Comment
Diffusion couple HBS-USTC-DC1
a b
HBS-USTC-DC2
a
1697
1712±20
161
174
b HBS-BGI-DC3
a
1593
1608±20
189
209
b HBS-BGI-DC4
a
1643
1658±20
172
187
b HBS-BGI-DC5
a
1813
1828±20
117
148
b HBS-USTC-DC6
a
1748
1763±20
144
163
b HBS-USTC-DC7
a
1831
1846±20
113
122
b
Elevated H2Ot
Hydration HBS-USTC-Hy1
1818
1833±20
149
159
0.03
0.03/0.06
HBS-USTC-Hy2
1747
1762±20
202
212
0.03
0.03/0.06
HBS-USTC-Hy3
1686
1701±20
157
168
0.03
0.03/0.06
HBS-USTC-Hy4
1666
1681±20
232
237
0.03
0.03/0.06
a
Each diffusion couple experiment contains two diffusion couples with different water contents.
b
Effective experimental duration at the corrected temperature (see text for details).
c
FTIR-determined water contents of the starting glasses.
d
FTIR-determined water contents in the post-experiment samples.
profile.
For diffusion couple runs, the numbers represent water contents at the two flat regions of the diffusion
For hydration runs, the numbers are the water content at the flat region (sample interior) and that at the sample-capsule interface, respectively. 29
Table 3 Definition of symbols Notation C X x t D P T K
c F H z R
Definition
Unit/Value
Weight percentage Mole fraction Distance Time Diffusivity Pressure Temperature Equilibrium constant Electrical conductivity Ion concentration Faraday’s constant Haven ratio Valence Gas constant
wt% m s m2/s Pa K S/m mol/m3 96485 C/mol
8.3145 J/(mol·K)
30
Table 4 OH and H2Om diffusivities extracted from fitting using the modified speciation model Run # Diffusion couple HBS-USTC-DC1 HBS-USTC-DC2 HBS-BGI-DC4 HBS-BGI-DC5 HBS-USTC-DC6 HBS-USTC-DC7
logDOH
logDH2Om
R2
1846
0.12 0.10 0.11 0.17 0.12 0.10 0.10 0.12 0.14 0.14
–9.64±0.01 –9.58±0.01 –9.53±0.01 –9.50±0.01 –9.73±0.01 –9.41±0.01 –9.39±0.01 –9.31±0.01 –9.33±0.01 –9.07±0.01
–8.73±0.01 –8.58±0.01 –8.57±0.01 –8.73±0.01 –8.81±0.01 –8.39±0.01 –8.40±0.01 –8.39±0.01 –8.48±0.01 –8.21±0.01
0.9997 0.9998 0.9999 0.9999 0.9999 0.9998 0.9997 0.9997 0.9998 0.9995
1818 1747 1686 1666
– – – –
–9.12±0.02 –9.36±0.03 –9.59±0.05 –9.64±0.04
– – – –
0.9981 0.9957 0.9869 0.9944
T (K) a b a b a a b a b b
Hydration HBS-USTC-Hy1 HBS-USTC-Hy2 HBS-USTC-Hy3 HBS-USTC-Hy4
1708 1712 1658 1828 1763
Best DOH/DH2Om
Errors are given at 2 level.
31
Figure Captions
Fig. 1. Diffusion profile for sample HBS-USTC-DC6a measured by FTIR and fitted by different models.
(a) Profile and fits; and (b) Inferred H2Ot diffusivity as a
function of the mole fraction of H2Ot.
Erfc: error function (constant diffusivity);
exponential: DH2Ot = D0eaX; speciation-based model: DH2Om = D0eaX and DOH = 0. Best fit was obtained using the modified speciation-based diffusivity model: DH2Om = constant and DOH = constant, and the inferred H2Ot-dependent DH2Ot is in good agreement with the result from Boltzmann-Matano analysis.
Fig. 2. Diffusivities of H2Om and OH in basaltic melt at 1 GPa extracted from least squares fitting of experimental diffusion profiles (Table 4).
Open diamonds are OH
diffusivity (≈ H2Ot diffusivity) from error function fitting of hydration profiles.
For
diffusion couple profiles fitted by the modified speciation-based model, H2Om diffusivity and OH diffusivity are shown in solid circles and solid diamonds, respectively.
Straight lines are Arrhenius fits corresponding to equations (5) and (6).
Error bars are shown at 2 level while the vertical error bars are within the size of symbols.
Fig. 3. Absolute (solid curves) and relative contribution (dashed curves) of OH and H2Om to H2Ot diffusivity in basaltic melt at 1846 K and 1 GPa.
Fig. 4. (a) Total water diffusivity at 1 wt% H2Ot and 1 GPa in silicate melts, compared with calculations using the model (dashed lines; their eqn. 26) in Zhang and Ni (2010). 32
(b) Total water diffusivity at 0.2 wt% H2Ot and 1 GPa in silicate melts.
Data sources:
N(13): Ni et al., (2013); N&Z(08): Ni and Zhang (2008); B(04): Behrens et al. (2004); Z&S(91): Zhang and Stolper (1991).
Fig. 5. Comparison of the diffusivities of hydrous species with Na diffusivity in nominally anhydrous basaltic and andesitic melts. melt are from Ni et al. (2013).
Water diffusivities in andesitic
Na diffusivity from Lowry et al. (1982) is similar in
the two melts.
Fig. 6. Contribution of OH to electrical conductivity relative to the contribution of Na in two hydrous silicate melts, based on the Nernst-Einstein equation using the diffusivity data from Lowry et al. (1982), Ni et al. (2013) and the present study.
Fig. 7. Bubble growth in basaltic melt and rhyolitic melt with 2 wt% H2O at 1573 K and 50 MPa using the model of Liu and Zhang (2000). this study and Ni and Zhang (2008).
Water diffusivity is based on
Water solubility and melt viscosity are based
on Zhang et al. (2007) and Hui and Zhang (2007), respectively. radius is set to be 1 m.
33
The initial bubble
Figure1
0.04
(a) HBS-USTC-DC6a 1763 K, 163 s
1.5
0.02
1 FTIR erfc exponential speciation modified speciation
0.01
0 -1500
0.5
0 -1000
-500
0
500
1000
1500
x (μm) CH2Ot (wt%) 2000
0
0.5
1
1.5
2
(b) HBS-USTC-DC6a 1763 K, 163 s
DH2Ot (μm2/s)
1500
1000
erfc exponential speciation modified speciation Boltzmann-Matano method
500
0
0
0.01
0.02
XH2Ot
0.03
0.04
CH2Ot (wt%)
XH2Ot
0.03
2
Figure2
T (°C) -8
1600
1550
1500
1450
1400
1350
logD (D in m2/s)
-8.5
DH2Om -9
DOH
-9.5
-10 0.54
0.56
0.58
1000/T (T in K)
0.6
0.62
Figure3
1
2500 OH
D (μm2/s)
2000
0.8
D H2Ot
1500
D H2O
0.6
X X m/d d · m
0.4
1000
DOH·dXOH/2dX
500 0
0.2
H 2O m
0
0.5
1
H2Ot (wt%)
1.5
2
0
Contribution of hydrous species
1846 K, 1 GPa
Figure4
T (°C) -8
1700
1600
1500
1400
1300
1200
logDH2Ot (D in m2/s)
(a) MO
-9 Bas alt A nd
RB Z&S (91
(thi s st udy
esit e
-10
B(0 4
)
1 wt% H2Ot ~ 1 GPa )
)
Dac ite B (
04) Rhyo lite N &Z(0 8)
-11 0.5
0.55
0.6
0.65
0.7
1000/T (T in K) T (°C) -8
1700
1600
1500
1400
1300
1200
logDH2Ot (D in m2/s)
(b) -9
Bas alt
(thi s st udy
MOR B Z& S (91)
)
-10 A nd
esite
N (1
3)
Daci te B( 04) Rhyol ite N& Z(08)
-11
-12 0.5
0.2 wt% H2Ot ~ 1 GPa
0.55
0.6
1000/T (T in K)
0.65
0.7
Figure5
DH2Om
T (°C) -8.5
1700
1600
1500
1400
1300
logD (D in m2/s)
-9
DNa
DOH -9.5 Basalt
-10
DOH -10.5
-11 0.5
Andesite
0.55
0.6
1000/T (T in K)
0.65
Figure6
10
6w Basalt (
1
σOH/σNa
2w Basalt (
Andesite
0.1
) t% H 2O t ) t% H 2O t
O) (6 wt% H2 t
(2 wt% Andesite
0.01 1200
1300
1400
T (°C)
H2Ot)
1500
1600
Figure7
bubble radius (μm)
300
1573 K 50 MPa 2 wt% H2O
200
basalt
100
rhyolite
0
0
500
1000
t (s)
1500
2000