You are on page 1of 14

Contributions to Mineralogy and Petrology (2019) 174:1

https://doi.org/10.1007/s00410-018-1536-4

ORIGINAL PAPER

Methane-bearing fluids in the upper mantle: an experimental


approach
Vladimir Matjuschkin1 · Alan B. Woodland1 · Gregory M. Yaxley2

Received: 15 February 2018 / Accepted: 16 November 2018


© Springer-Verlag GmbH Germany, part of Springer Nature 2018

Abstract
The main obstacle to understanding of the geological role of reduced, ­CH4-bearing fluids is the absence of a reliable experi-
mental technique applicable to solid-media high-pressure apparatuses, allowing their observation and direct characterisation
under laboratory conditions. In this study, we describe the main pitfalls of earlier designs and technical aspects related to
achievement of strongly reduced oxygen fugacity (fO2) conditions (i.e., Fe–FeO, IW) and maintenance of a constant fluid
equilibrium during an experiment. We describe a new triple-capsule design made of an Au outer capsule with an Au-inner
capsule containing a metal/metal oxide oxygen buffer and water, as well as an inner olivine container filled with a harzburgitic
sample material and Ir powder that serves as a redox sensor. The bottom of the outer capsule is covered with a solid fluid
source (e.g., stearic acid). The outer capsule is surrounded by a polycrystalline ­CaF2 pressure medium to minimise ­H2-loss
from the assembly. Application of this design is limited to temperatures below the melting temperature of Au, which is pres-
sure dependent. Metals other than Au can lead to fluid disequilibrium triggered by a dehydrogenation and carbonation of the
methane. Test experiments were carried out at 5 GPa, temperatures < 1300 °C, at Mo–MoO2 and Fe–FeO buffer conditions.
IrFe alloy sensors demonstrate successful achievement and maintenance of reduced fluid environment at ∆logfO2 ≈ IW + 0.5.
The fluid phase was trapped in numerous inclusions within the olivine sample container. Raman spectra reveal that the fluid
consists mainly of C ­ H4, along with small amounts of higher hydrocarbons like C ­ 2H6. No water was detected, but H ­ 2 was
found to be present in fluid and incorporated into the olivine structure. Our results are inconsistent with published fluid
speciation models that predict significant ­H2O contents at these fO2 conditions. It is also apparent that fluids with significant
­CH4 contents are likely to be stable under the conditions recorded by some mantle samples.

Keywords  Reduced fluid · Methane · Upper mantle · High pressure · Experiments · Oxygen fugacity buffer · Graphite
saturation

Introduction

The role of fluids is fundamental for most geological pro-


cesses. For instance, they control the solidus temperature of
rocks as well as the physical parameters of subsequent melts,
Communicated by Jochen Hoefs.
such as viscosity and density. Fluid components of mag-
Electronic supplementary material  The online version of this mas also affect their crystallisation sequence and eruption
article (https​://doi.org/10.1007/s0041​0-018-1536-4) contains behaviour (Blundy and Cashman 2008; Blundy et al. 2010;
supplementary material, which is available to authorised users. Melekhova et al. 2013; Zellmer et al. 2015). The extreme
mobility and high redox potential of fluids containing C,
* Vladimir Matjuschkin
v.matjuschkin@em.uni‑frankfurt.de O, and H (and also S) means that they are important meta-
somatic agents and can exert a strong influence on oxida-
1
Institut für Geowissenschaften, Goethe-Universität Frankfurt tion–reduction reactions, largely contributing to the forma-
am Main, Altenhöferallee 1, 60438 Frankfurt am Main, tion of redox heterogeneities within the Earth (Arculus 1985;
Germany
Evans and Tomkins 2011; Zellmer et al. 2015; Matjuschkin
2
Research School of Earth Sciences, The Australian National et al. 2016; Uenver-Thiele et al. 2017). Clearly, the nature
University, Canberra, ACT​ 2601, Australia

13
Vol.:(0123456789)
1  Page 2 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

and dynamics of magmatic fluids are important parameters composition or even with the humidity in the laboratory
to constrain. Their composition and thermodynamic proper- (e.g., Matjuschkin et al. 2015). Intrinsic fO2 results from the
ties are all strong functions of oxygen fugacity (fO2) (French hydrogen exchange between the sample and the surround-
1966; Zhang and Duan 2010) and pressure, owing to their ing parts of the apparatus, which is a strong function of the
high compressibility compared to solid phases. This makes experiment temperature. Intrinsically controlled water-rich
modelling fluid compositions very challenging. samples often suffer from water loss, followed by strong oxi-
Information about the behaviour of deep-seated mag- dation of the sample material and then later by a reduction to
matic fluids or those existing in the Earth’s upper mantle is the ‘intrinsic value’. This fO2 value generally varies between
mostly derived from experimental studies. Thermodynamic − 0.5 and + 3 log units relative to the synthetic fayalite–mag-
calculations rely on experimental calibrations, which are netite–quartz oxygen buffer (i.e., FMQ − 0.5 to FMQ + 3) for
difficult to produce at geologically relevant high pressures a piston–cylinder or belt apparatus at 3 GPa and ~ 1000 to
and temperatures, where fO2 must also be controlled and 1200 °C (unpub. Data; also see Jakobsson 2012; Jakobsson
well monitored. At low pressures (< 0.5 GPa), externally and et al. 2014a, b; Stamper et al. 2014). Intrinsic fO2 control is
internally heated gas-pressure vessels can be employed with often employed in combination with relatively “dry” sample
fO2 controlled by directly changing the hydrogen fugacity materials and is not suited for studying fluid-rich samples at
(fH2) of the gas-pressure medium. Such control is not pos- any fO2, since equilibrium conditions cannot be guaranteed
sible at higher pressures using solid-media apparatuses (e.g., nor subsequently demonstrated.
piston–cylinder 0.5–3 GPa, belt apparatus 3–6.5 GPa, and
multi-anvil press > 7 GPa), and fluid composition is usu- Existing fH2‑buffering techniques and their
ally fixed following the double-capsule buffering technique limitations
developed by Eugster (1957). This approach is less accurate
and reliable (Matjuschkin et al. 2015), particularly at low fO2 To solve the fO2 problem for high-pressure experiments,
and a number of pitfalls need to be avoided when assembling Eugster (1957) introduced the double-capsule technique,
and running a successful experiment. which relies on the principle of hydrogen exchange between
In this contribution, we first focus on common problems a solid oxygen buffer assemblage (e.g., Mo–MoO2) and the
related to the control and monitoring of fO2 and fluid spe- sample materials that are separated by a metal membrane
ciation, especially under strongly reducing conditions (i.e., that is permeable to ­H2. When fH2 equilibrium between
around the iron–wüstite, IW, oxygen reference buffer), where buffer and sample is achieved, a sample containing a pure
­H2 and ­CH4 are expected to be significant fluid components. ­H2O fluid (aH2O = 1) also attains the oxygen fugacity of the
We then go on to describe a new, modified capsule design metal buffer through the reaction:
that aims to solve or at least minimise the main drawbacks
and problems of existing designs. All of our test experi- H2 O = H2 + 0.5O2 , (1)
ments were performed with a harzburgitic bulk composition as long as the fluid phase in the sample and buffer assem-
at pressures of 3–5 GPa and temperatures up to 1300 °C, in blage is pure water. However, at high pressures and tem-
a belt apparatus. However, such a design can be applied to peratures, the fluid may contain significant dissolved com-
fluid-bearing experiments in any other solid-media apparatus ponents, causing a systematic shift in the desired fO2 due to
having a cell volume of > 200 mm3 (i.e., piston–cylinder, a deviation from unit water activity (Sisson et al. 2005). In
large volume multi-anvil press). addition, as the sample equilibrates at the new fO2 imposed
by the buffer assemblage it will either gain or lose water
depending on whether it underwent relative reduction or oxi-
Important aspects related to the control dation, respectively (i.e., the forward or reverse of reaction
and monitoring of fO2 1, respectively). This effect is important to consider when
determining the relative size of the buffer employed and the
“Intrinsic” fO2 control amount of added ­H2O.
The original double-capsule designs and their modifica-
In absence of a fO2 buffer, the sample of a high-pressure tions (e.g., Eugster 1957; Jakobsson et al. 2014a, b; Moore
experiment is essentially ‘unbuffered’. Such an experiment et al. 2008; Sokol et al. 2009; Tiraboschi et al. 2016) have
approaches the fO2 of the apparatus as dictated by its con- since undergone several improvements. Another recent
stituent parts (e.g., furnace, pressure cell materials, etc.) modification was introduced by Matjuschkin et al. (2015)
and chemical environment (e.g., fH2), and this is referred and consists of a triple-capsule arrangement with an exter-
to as ‘intrinsic fO2 control’ or ‘intrinsically buffered fO2’. nal Pyrex sleeve, which together help to (1) reduce ­H2 loss
This term is very misleading, since no actual control is and promote the early achievement of fO2 equilibrium in the
provided and fO2 can vary with run duration, sample bulk sample (2) reduce the temperature gradient in the assembly

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 3 of 14  1

owing to the large amount of H ­ 2O present in the outer cap- are also available for producing fluids at carbon saturation
sule and (3) prevent chemical interaction between the sample (e.g., tartaric acid, stearic acid, and lauric acid). In practice,
and the buffer assemblage. solid-organic materials require ~ 2 h equilibration time at
In principle, the designs of Eugster (1957) and T > 1100 °C (Sokol et al. 2017) and reach slightly higher fO2
Matjuschkin et al. (2015) can be used in combination with than expected, since theoretical calculations for estimating
a variety of available redox buffers. In the following text, fluid composition often ignore the presence of hydrogen.
we will demonstrate that these designs are only suited for For instance, during heating stearic acid (­ C18H36O2) breaks
experiments with pure ­H2O as the fluid phase or for rela- down to ­8CH4 + 2H2O + 10C with a ­CH4-to-H2O ratio of 4:1,
tively oxidised, ­H2O–CO2 fluids lying in the fO2 region at or which is used to calculate the equilibrium fO2 of the experi-
above the ‘water maximum’ (see Woermann and Rosenhauer ment. However, amount of ­CH4 is actually somewhat lower,
1985). because part of it reacts to C + 2H2 to produce the required
fH2 dictated by the buffer.
Solid and fluid fO2 buffers
Monitoring of fO2
A variety of materials can be employed as redox buffer
assemblages. Metal–metal oxide pairs together with ­H2O are Tracking and monitoring fO2 can be achieved by various
generally favoured due to the rapidity with which the buffer means. It can be calculated from mineral compositions (e.g.,
equilibrium and the appropriate fH2 can be established soon QUILF, Andersen et al. 1993; spinel–olivine–orthopyroxene
after the onset of the experiment. In addition, such assem- assemblages; Wood 1990; Ballhaus et al. 1991) mineral-
blages also have the advantage of a high buffering capac- melt equilibria (e.g., ol-melt, Matzen et al. 2011) or from
ity per unit volume. Common buffers that cover a range in melt composition (e.g., ­Fe3+/∑Fe Borisov et al. 2015; Kress
fO2 include Re–ReO2, Ni–NiO, Co–CoO, and Mo–MoO2 as and Carmichael 1991). However, many of these approaches
well as the Fe-bearing pairs F­ e2O3–Fe3O4 (HM), F ­ e3O4–FeO are not applicable as redox sensors in experiments, particu-
(WM) and Fe–FeO (IW). Depending on whether the sample larly under reduced conditions, due to their relatively low
fO2 is to be significantly shifted during the experiment, the precision.
relative proportion of metal to oxide in the buffer assem- A number of different metals coexisting with oxides or
blage should be adjusted to improve its buffering capacity silicates have been calibrated as redox sensors. For example,
(i.e., a high M/MO for achieving reduction of the sample and (Ni,Mn)O/Ni (Pownceby and O’Neill 1995) and (Co,Mn)O/
the opposite if a net oxidation is desired). This is particularly Co (Pownceby and O’Neill 2000), along with Ni–Pd/NiO
the case when the target fO2 significantly deviates from the (Pownceby and O’Neill 1994a, b) and Co–Pd (Taylor et al.
intrinsic fO2 of the pressure cell and apparatus. The average 1992), can deliver precise results over different ranges of
run time required to effect a significant change in ­Fe3+/∑Fe fO2. However, these may be difficult to use if they are to be
or ­S6+/∑S of the sample is about 12–24 h, assuming that the in physical contact with samples having natural rock com-
mass of the sample is relatively small compared to that of the positions, since Ni, Co, and Mn can be readily incorporated
buffer [< 1/3 of mbuffer, (Matjuschkin et al. 2015)]. in many geologically relevant phases. It is known that some
Metal–metal oxide buffers can be physically separated experimentalists separate sensors by placing them next to
from the sample in their own capsule (e.g., Eugster 1957; the capsule assembly. Such sensors are physically separated
Matjuschkin et al. 2015; Moore et al. 2008) or be in direct from the sample and the capsule by zirconia tubes or other
contact (Matveev et al. 1997). In the latter case, potential materials. In this case, sensor readings might be interpreted,
reactions with the sample must be considered, such as the as “intrinsic fO2” values of the apparatus, however, do not
undesired formation of mineral solid solutions or dissolution represent the fO2 inside the sample (see further effects of
in a melt phase. pressure medium material, hydrogen loss, water loss, and
Compared to solid-state buffers, fluids (e.g., COH flu- Fe loss).
ids) exhibit much lower buffering capacity. However, the Redox sensors based upon Fe-bearing noble metal alloys
choice of fluid components can aid in the rapid attainment offer a way around the “sample contamination” problem,
of the target fO2. For example, ­CO2–H2O components can since solid solution of noble metals is limited in minerals
be accurately added as hydroxides, carbonates, and oxa- under the majority of redox conditions of geologic interest.
lates (e.g., Holloway et al. 1968). These components must The most commonly used alloys are Pt–Fe (Jamieson et al.
be prepared and stored under relatively dark, cold, and 1992), Pd–Fe (Borisov and Palme 2000), and Ir–Fe (Wood-
dry conditions to avoid adsorption of C ­ O2 (hydroxides, land and O’Neill 1997). More complex ternary alloys of
oxalates) or oxidation (oxalates) (Boldyrev 2002a, b). A Au–Pd–Fe (Barr and Grove 2010) have also be employed.
variety of other COH-bearing substances that are solid at The use of Ir–Fe alloys has several advantages over those
room temperature, but break down at elevated temperature, with Pt or Pd. The first is the stability of face-centred cubic

13
1  Page 4 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

(fcc) γ–Ir–Fe across the entire binary over a large tempera- is detrimental to a successful experiment. Hydrogen per-
ture range, without intervening intermediate phases (at least meabilities vary widely between the different noble met-
at 1 bar, Cahn 1991). Second, the solubility of Ag and Au in als with Pd > Ag70Pd30 > Ag80Pd20 ≥ Pt > > Ag > Au (Chou
Ir is very low (Cahn 1991), which is of importance, because 1986; Matjuschkin et al. 2015).
Ag and Au are potential capsule materials for high-pressure Au exhibits the lowest ­H2 permeability, but nevertheless
experiments. Finally, low carbon solubility in γ–Ir–Fe allows has often been successfully used as a hydrogen membrane
their use in reduced carbon-bearing systems. The activity at > 800 °C (Matjuschkin et al. 2016) or even at lower tem-
coefficient of Fe in Ir–Fe alloys is also well constrained, peratures in long-duration cold-seal experiments (Mutch
allowing the activity of Fe in the experimental assem- et al. 2016). Hence, Au can be used in reduced experiments,
blage to be accurately calculated from microprobe analysis where high volume of hydrogen transfer from buffer to sam-
(Woodland and O’Neill 1997). This result can subsequently ple is required. The fact that Au can be easily welded means
be applied to Fe-bearing redox equilibria to derive the fO2 that failure due to catastrophic fluid loss is rare. The opera-
during the experiment. tional temperature range is from 650 to 1250 °C at ~ 3.5
GPa (Akella and Kennedy 1971). Below 650 °C, hydrogen
diffusion through Au is so slow that the metal effectively
Reduced environments at high pressures becomes a permeability barrier (Chou 1986). Somewhat,
higher temperatures can be reached at higher pressures as the
Achieving and maintaining reducing conditions (i.e., at or melting point of Au increases (Akella and Kennedy 1971).
near IW) in high-pressure experiments present additional Another consideration for reduced experiments is the
challenges for a number of reasons. One important aspect is potential loss of Fe to the capsule material during the experi-
the high fH2 that must be maintained in the capsule, but the ment. This can not only modify the bulk composition of the
intrinsic fO2 of the apparatus normally defines a much lower sample, but also cause oxidation via
fH2. This leads to a strong fH2 gradient between the sample
FeO + M = MFe + 0.5O2
and the assembly outside of the capsule and can facilitate , (2)
sample capsule capsule sample
rapid ­H2 loss from the sample. It is essential that H
­ 2 loss is
minimised in order for a constant fO2 to be maintained for where M is the capsule metal. At fO2 above the ‘water
the duration of the experiment (Matveev et al. 1997). This maximum’ (i.e., graphite saturation, Woermann and Rosen-
requires careful consideration of the type of materials used hauer 1985), Fe incorporation in AuPd or Pt ranges from 0
for the capsule as well as for the parts of the pressure cell to 30 mol % as a function of fO2, and this limited amount
directly around the capsule. Further complications of experi- may not be viewed as a serious problem (e.g., Borisov and
ments at low fO2 include the potential for substantial Fe loss Palme 2000). Fe loss can be minimised by pre-saturating the
from the sample to the capsule. In C-bearing experiments, capsule materials with Fe (which might make welding more
fluids will be ­CH4 rich, meaning that speciation is more difficult) or using Au capsules. However, at more reducing
complex, with H being distributed between multiple spe- conditions, significant Fe loss can be expected to occur even
cies, potentially including higher order hydrocarbons. There with pure Au metal (Cahn 1991; Borisov and Palme 2000;
is also the possibility that unwanted fluid–metal reactions Ratajeski and Sisson 1999). Under such reduced conditions,
may compromise fO2 buffering in an experiment. We now Pd and Pt-bearing alloys can incorporate substantial amounts
address these potential problems in turn and provide recom- of Fe and their use for capsules should be avoided. The lack
mendations on how to solve or at least mitigate these effects. of significant alloying between Fe and Ag (Massalski et al.
1986) makes the later an ideal capsule material as long as the
Choice of capsule material relatively low maximum working temperature is compatible
with the required experimental conditions. Otherwise, Au
Aside from the normal constraints of isolating the sam- can be employed, keeping in mind that oxidation due to Fe
ple from its surroundings, fluid-bearing experiments loss must be compensated for in one way or another (i.e., by
necessitate the use of metal capsules. As demonstrated an internal buffer or fluid phase, see below).
by Matjuschkin et al. (2015), at high pressure and tem- Working with C-bearing fluids adds an additional consid-
peratures, hydrogen leakage from the capsule assembly is eration to the choice of capsule material, since some metals
inevitable, since all noble metals are permeable to hydro- can induce chemisorption and carbonisation reactions. It is
gen to some degree or other at > 650 °C (also see Chou known that reduced, ­CH4 rich, COH fluids are unstable in
1986; Truckenbrodt and Johannes 1999). This is useful in the presence of Pt, Pd, or AuPd alloys at T > 600 °C (and
double and triple-capsule assemblies, as it allows rapid other metals as Ni, Co, Fe, Mo, Ir, Rh, Ru, etc.) (Concha
equilibration of fH2 between the buffer and sample assem- et al. 1984; Hanley et al. 1991; Rostrup-Nielsen 1993; Soly-
blages. On the other hand, ­H2 loss from the outer capsule mosi and Erdohelyi 1980; Vannice 1977; Christensen et al.

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 5 of 14  1

2006). For instance, contact of C


­ H4 with Pd causes a step- assemblage. Without a metal–metal oxide buffer, the fluid
wise dehydrogenation reaction (Rostrup-Nielsen 1993): composition shifts to more oxidised field towards the water
maximum (Matveev et al. 1997). Other materials, such as
CHfluid
4
+ Pd → ⋯ → C − Pd + 2H2 , (3) ­Al4C3–Al(OH)3, have also been used (Taylor and Green
or 1987; Taylor and Foley 1989), but this combination yields
CH4 − Pd → CH3 − Pd + 1∕2H2 → CH2 − Pd + H2 → CH − Pd + 3∕2H2 → C − Pd + 2H2 . (4)

Similar chain reactions are expected for other hydrocar- ­Al2O3 that can contaminate a sample. It also produces essen-
bons via (Rostrup-Nielsen 1993) tially pure C
­ H4 that must then further react to form an equi-
librium fluid composition containing H ­ 2 and H­ 2O, placing
m
Cn Hm → ⋯ → nC + H. (5) an even higher demand on the buffer assemblage for ­H2 than
2 2
substances like stearic or lauric acid.
Even more oxidised CO-bearing fluids are at risk of ‘car-
bon loss’ at pressures < 3 GPa, unlike C
­ O2-fluid that remains
nearly unaffected by contact with Pd or Pt (Solymosi and The importance of the material surrounding
Erdohelyi 1980; Evans et al. 2006). We suspect that this may the capsule
be the reason for the CO-fluid component being so difficult
to detect in experimental run products. The dehydrogenation The surrounding pressure medium (‘spacer’, ‘packing mate-
of hydrocarbons is a clear problem and makes the predic- rial’) is a critical part of the experimental assembly. Ideally,
tion of and the maintenance of fO2 very challenging. Short- it creates a chemically closed environment around the outer
duration experiments (i.e., 1–3 h) do not provide a solution, capsule and protects it from physical damage. For instance,
since fluid–mineral equilibrium cannot be assured, but long soft metal capsules (e.g., Au, ­Au90Pd10, and Ag) placed
durations can lead to severe changes in the fluid composi- in a material with a large porosity (e.g., pressed powders,
tion. Graphite-lined capsules can be used to slow down the crushable alumina, or MgO) can experience strong degra-
‘carbon loss’ reaction, but do not eliminate it. The references dation or failure within 24–48 h (Fig. 1a, b). This problem
given above provide numerous examples of this effect in can be avoided using less porous materials or glasses (e.g.,
industrial processes. ‘Pyrex’ glass, soda–lime–silica glasses, and polycrystalline
In contrast to Pt, Pd, or AuPd alloys, Au does not inter- ­CaF2). Harder capsule materials (Pt and Au–Pd alloys with
act with the fluid and can preserve fluid composition much Pd > 20 mol %) are more resistant to failure (Rosenbaum and
longer, provided that fH2 is externally buffered and main- Slagel 1995), but are not suitable for low fO2 experiments,
tained (this study at high-pressures; also see Kratzer et al. as previously described.
1996). Thus, it is advisable for reduced experiments contain- Apart from protecting the capsule, the surrounding pres-
ing methane to only use Au capsules. sure medium is also very important for slowing down ­H2 loss
and helping to maintain fluid inside the sample capsule. Poor
Producing reduced COH fluids hydrogen isolators are porous, crushable materials, powders,
vesicle-rich glasses, or very soft materials (e.g., commonly
Under reduced conditions, below the ‘water maximum’, used crushable alumina and MgO or NaCl). Such materials
­CH4 and ­H2O are the dominant fluid species in a COH fluid. lead to uncontrolled H ­ 2 exchange between the capsule and
Such compositions can be added in the form of solid-organic external parts of the assembly (Fig. 1b) and press (Freda
compounds with low thermal stabilities, which helps to ini- et al. 2001). Note that industrial ceramic (­ Al2O3) is a good
tially establish a low fO2 fluid without exhausting the capac- hydrogen isolator, but is useless as a pressure medium, since
ity of the buffer assemblage. For instance, anthracene or it cannot provide a hydrostatic pressure distribution around
docosane ­(CH4/H2O ≈ 1:0), stearic acid (4:1), or lauric acid the capsule (Matjuschkin et al. 2015; Serra et al. 2005).
(5:2) can be used to rapidly achieve logfO2 levels from ~ IW Thus, the physical properties and quality of the material
to ~ IW + 3, respectively (calculated from Belonoshko and determine its suitability as a pressure medium, particularly
Saxena 1992; Zhang and Duan 2010). The breakdown of in fluid-rich experiments under reducing conditions.
these compounds also produces variable amounts of car- OH-bearing materials (i.e., unfired pyrophyllite or talc)
bon, i.e., one mole of stearic acid yields ten moles of C: are often used to prevent the water loss by increasing the
­C18H36O2 = 8CH4 + 2H2O + 10C. The “missing” ­H2 nec- overall ‘intrinsic fH2’ of the assembly (Freda et al. 2001;
essary for an equilibrium COH fluid at fO2 of IW will be Truckenbrodt et al. 1997). Such materials are typically
generated by a combination of the partial breakdown of used for short-duration experiments, since these materi-
­CH4 ­(CH4 = C + 2H2) and H ­ 2 provided by an internal buffer als dehydrate over time, effectively changing the external

13
1  Page 6 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

Fig. 1  a Example of metal capsule degradation as illustrated by two the experiment. b Focused hydrogen diffusion through the bottom of
(soft) Au capsules in a hard, porous pressure medium material com- an outer gold capsule caused hydrogen corrosion (darker dots within
posed of a crushable ­Al2O3 from a piston–cylinder experiment at the Au). This was triggered using a porous crushable MgO disk at
0.7 GPa, and 850  °C, with a 72-h run duration (photo, courtesy of the base of the assembly, whilst other walls were protected by a
Richard Brooker). The coloured dots in ­Al2O3 is Au from the capsule solid Pyrex-glass sleeve. Assembly from an experiment at 1 GPa and
material that was pressed into the pressure medium material during 1000 °C for 24 h (after Matjuschkin et al. 2015)

fH2 during the course of the experiment. This may lead to discordant fO2 values registered by the two sensors, as well
inconsistent results within a series of experiments con- as different results as a function of experiment duration.
ducted at different temperatures or durations and is unsuit- Upon opening the capsules, ~ 3.4-µL H ­ 2O (of 4.5 µL)
able for experiments, where exact fO2 control is desired. was emitted, indicating substantial conservation of the ini-
Pyrex glass helps to significantly retard ­H 2 loss, but tial water content. Analysis of the sensor alloys by EPMA
cannot guarantee the maintenance of constant fH 2 in (see supplementary information) revealed homogeneous
experiments longer than ~ 24 h (Matjuschkin et al. 2015). compositions with no obvious zonation. Compositions of
Therefore, fluid-buffered experiments with low buffering ­Co25Pd75 and N ­ i42Pd58 produced in the 12-h experiment
capacity always run at the risk of being exhausted unless and are almost identical those obtained from alloys in the
an additional buffer is present. 24-h experiment ­(Co27Pd73, ­Ni46Pd54) (Fig. 2). Thus, no sig-
As an alternative to Pyrex, we tested the use of dense nificant shift in internal fO2 occurred over a 24-h period,
polycrystalline ­CaF2 (not pressed powder, C.Giese GmbH indicating that dense polycrystalline C­ aF2 is an ideal pres-
& Co.KG, Idar-Oberstein, Germany) as the pressure sure cell material for fluid-bearing experiments requiring a
medium surrounding the capsule, as well for the entire constant fH2.
pressure cell (Brey et al. 1990). As the ability of ­CaF2 to
inhibit ­H2 migration is currently unknown, we performed
two experiments with our assembly following the approach The new experimental design
of Matjuschkin et al. (2015). Two 3-mm Au capsules were
prepared, each containing C ­ o28Pd72 and ­Ni64Pd36, as fO2 Taking the above considerations into account, we have
sensors (Pownceby and O’Neill 1994a; Taylor and Foley developed an experimental design that allows us to reach
1989) together with 4.5-µL ­H 2O. The two alloys were and maintain reducing conditions for sample bulk composi-
separated from each other and from the capsule walls by tions relevant to the Earth’s mantle. The setup was conceived
­Al2O3 powder to avoid unwanted reaction. The capsules for a belt apparatus, but can just as well be employed in
were welded shut and inserted into C ­ aF2 pressure cells, a piston–cylinder apparatus or a large volume multi-anvil
where they were run at 3 GPa and 1000 °C for 12 and press. In the latter case, a 25-mm octahedron pressure cell
­ 2 loss is low, water will remain in
24 h, respectively. If H is necessary to accommodate the capsule. The design com-
the capsule and both sensors will equilibrate/react to simi- prises three major parts: (1) a 4.4-mm diameter Au outer
lar fO2 values. If progressive ­H2 loss occurs, water will capsule that encloses; (2) a small fO2-buffer capsule; and (3)
also be lost, leading to continuous fH2 disequilibrium and an olivine capsule that contains the sample mixture (Fig. 3).

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 7 of 14  1

-7
3 GPa, 1000°C For the fO2 buffer (IW, Mo–MoO2), only the metal compo-
NiPd sensor nent (Fe or Mo) is loaded together with 3.5-µL H ­ 2O into a
CoPd sensor IW +6
3.0-mm Au capsule and welded shut. The oxide component
-8 forms at the onset of the experiment and starting with just
IW +5
metal maximises the buffering capacity. The capsule is then
stable ƒH2 pressed from all sides to form a sphere. Round buffer cap-
logƒO2

-9 no further oxidation
sules are more stable during pressurization and remain cen-
ion
ibrat IW +4 tred, without tilting towards the walls of the outer capsule.
uil NNO buffer
eq
-10 The buffer capsule is placed into the outer capsule on top
of stearic acid powder that serves as the source of the COH
IW +3
fluid (Fig. 3). The use of Au avoids potential dehydrogena-
-11
0 10 20 30
tion reactions with the subsequent reduced, C ­ H4-rich fluid.
Run duration, hours The mass of stearic acid is 7 wt% of the total mass of the
sample, producing 4 wt% ­H2O + CH4 fluid and 3 wt% solid
Fig. 2  Test of hydrogen loss in an Au capsule surrounded by ­CaF2 C at the beginning of the run. The buffer and outer capsules
pressure medium material at 3 GPa and 1000 °C. Experimental pro- are physically separated by a filler powder that has the same
cedure is similar to that reported by Matjuschkin et  al. (2015). Col- bulk composition as the sample. For our test experiments,
oured symbols represent the readings of CoPd, NiPd redox sensors.
Comparing experiments with 12–24-h duration, the composition of
the sample comprised a finely ground powder of hand-
the sensor did not shift, indicating the absence of hydrogen exchange picked natural olivine and orthopyroxene (F7, Lazarov et al.
with the surrounding pressure medium material. Results are similar to 2009). The sample itself contained an additional 5 wt% of
those reported for Pyrex by Matjuschkin et al. (2015) ≤ 60-µm Ir powder to serve as a fO2 sensor. The sample is
packed into a 4-mm diameter olivine container (San Carlos)
that is inserted above the buffer into the outer Au capsule
(Fig. 3). The outer capsule is then welded shut and placed
into a polycrystalline ­CaF2 pressure cell.
Our test experiments were performed in a belt apparatus
at the Universität Frankfurt at 5 GPa and temperatures from
1100 to 1300 °C. The temperature was controlled using a
Type-B ­Pt70Rh30–Pt94Rh6 thermocouple, with ± 7 °C accu-
racy. Pressure calibration for the C ­ aF2 pressure cell was
sample reported by Brey et al. (1990). The experimental duration
+ varied from 12 to 24 h, depending on temperature (Table 1).
ƒO2 Experiments were quenched isobarically by shutting off the
sensor power supply to the graphite furnace. The recovered Au
capsule was embedded in epoxy and ground down until the
olivine
sample inside the olivine container was exposed. Paraffin
was used as a grinding medium, since liquid oils can pen-
ƒO2 Au etrate deep into graphite pockets and are difficult to remove.
buffer
sample
+H2O

Results
fluid source
Run products
CaF2
All experimental run products consist of re-equilibrated
olivine (ol) and orthopyroxene (opx) grains, graphite
1mm ­( C gr), together with a COH fluid (fl) (Table 1; Fig. 4a).
Analytical methods are described in the supplementary
materials. Olivine and orthopyroxene in the olivine con-
Fig. 3  New experimental setup (see text). Fluid source is thermally
tainer are compositionally uniform. The sample contact
unstable stearic or lauric acid. ­CaF2 is a solid polycrystalline mate-
rial enclosing the entire capsule assembly and separating it from the with the olivine container is sharp (Fig. 4a). In contrast,
thermocouple wire sample powder placed as filler next to the outer Au capsule

13
1  Page 8 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

Table 1  Overview of run conditions and experimental run products


Experiment P, GPa/T, °C Duration, h Buffer Mg# ol/opx XFe in Ir logfO2 sensor Sensor ∆ IW Fluid composition

1571 3/1150 24 IW 94/94 – – – CH4–H2, ­HCa


1572 5/1250 24 IW – – – – CH4–H2, ­HCa
1580 5/1250 24 IW 92/93 0.32 − 8.75 IW + 0.69 CH4–H2, ­HCa
1581 5/1280 18 IW 92/94 0.31 − 8.38 IW + 0.74 CH4–H2, ­HCa
1583 5/1200 20 IW 93/94 0.35 (14) − 9.53 (90) IW + 0.46 (90) CH4–H2, ­HCa
1585 5/1250 23 Mo–MoO2 93/94 0.30 − 8.66 IW + 0.78 CH4–H2, ­HCa
1586 5/1250 15 Mo–MoO2 93/94 0.30 − 8.18 IW + 0.73 CH4–H2, ­HCa
1587 5/1280 12 Mo–MoO2 93/94 0.34 − 8.61 IW + 0.50 CH4–H2, ­HCa
1611 5/1280 12 Mo–MoO2 94/94 0.33 − 8.45 IW + 0.47 CH4–H2, ­HCa

Ol and opx are re-equilibrated grains measured inside the ol-container (Mg#88) and are protected from the iron loss or buffer contamination,
respectively. Absence of some fO2 measurements is due to the sample loss during the preparation. Conservative uncertainty of fO2 measurements
is ± 0.05 log units and is a statistical value calculated from 5 to 20 analyses in each sample. High uncertainty of measurements in exp. 1583 are
due difficulty in analysing needle-like sensor aggregates (see text)
a
 Unspecified higher hydrocarbons, including ­C2H6

experienced strong Fe loss and resulted in up to 40%


FeO removal in olivine and up to 10% in orthopyroxene,
respectively (Fig. 4b).
Graphite is present in all experiments as a breakdown
product of stearic acid and appears either as needles and
clots filling the space between mineral grains or within
fluid inclusions. Some graphite was deposited on the walls
of the outer Au capsule or along cracks in the olivine con-
tainer indicating fluid circulation during the course of the
experiment.
Interestingly, no fluid was observed upon opening the
capsule. However, based upon textural evidence from
BSE images, a fluid phase was equally distributed within
the capsule (Fig. 4a). Considering that only 4 wt% fluid
equivalent was packed into the capsule, it appears that
the majority of the fluid has been trapped as inclusions
within the olivine container (Fig. 5a, b) or it occurred in
the pockets of graphite. Analysis of the fluid inclusions
by Raman spectroscopy reveals that they are composed of
methane and very minor ethane ­(C2H6) (Fig. 5c, d). Graph-
ite is present in some, but not all inclusions. No H­ 2O was
detected in any fluid inclusion. ­H2 was rarely detectable
as an additional fluid component from a Raman peak at
4155 cm−1 (Fig. 5d). In addition, some spectra of olivine
in non-confocal mode revealed the presence of ­H2 with a
broad peak position at 4119 cm−1 (Fig. 5d). This shift in
the peak position suggests that this ­H2 was incorporated
in the olivine structure, where it is better retained (Yang
et al. 2016). It is likely that H
­ 2 readily escapes from the
fluid during polishing (Yang et al. 2016) and when the
sample is placed under vacuum during EPMA investiga-
tion, thus lowering the probability of detecting this spe-
Fig. 4  a Experimental run products: the inner sample material with
cies in inclusions. The dominant presence of C ­ H 4 ± H 2
Ir sensor inside the olivine container in run 1581. b Effect of Fe loss
to the outer Au capsule, which affects the olivine composition and qualitatively testifies to the reducing conditions achieved
causes some transformation to orthopyroxene in our experiments.

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 9 of 14  1

Fig. 5  a, b Reflected polarised


light images of fluid inclusions (a) (b)
in the olivine container from
experiments 1585 and 1611,
respectively. c, d Representative gr
Raman spectra of fluid inclu- ol
sions in the olivine host. The
broad peak around that for C ­ H4 gr
at 2917 cm−1 is attributable to ol
other hydrocarbons, like C ­ 2H6
at 2948 cm−1 (Table 1) fluid
fluid
20µm 30µm

(c) 1586
C (d)
2722
arbitary arbitary
C
units units
2917 CH4 (L)
2917 CH (L)

2441 3065 H
CH (L) exp. 1580
C
2329
1358 N (air)
host olivine C 3247
H in host olivine:
C
host olivine 3065
1586 2722 CH (L)
- CH Mo-Mo Mo-Mo
C C 3247
stretching exp. 1586 exp. 1587
C
2869
1000 2000 3000 4000 cm-1 1000 2000 3000 4000 cm-1

The fO2 buffer (i.e., diffusion) through the buffer capsule (exp. 1585–1612,
Table 1) (Fig. 6c).
Results of test runs demonstrate that the use of appropriate
buffer components, in addition to their correct installation, Ir sensor and fO2 equilibrium
can significantly contribute to the overall success of an
experiment. Under strongly reducing conditions, employ- Initially, a small quantity of Ir powder was mixed with the
ing only metal is advantageous, since the significant ­H2 starting composition to sense the oxygen fugacity distri-
supply required is derived from equilibrium (1). This liber- bution inside the sample. At T ≥ 1250 °C, metal powder
ates oxygen that readily produces the MO buffer couple. reacted to a homogeneous alloy with platelets up to 30 µm
Fe-bearing buffers, like IW, tend to react even with the across ideally suitable for the electron-probe microanaly-
Au under such conditions. However, as long as both Fe sis (EPMA). Below 1250 °C, Fe alloyed more slowly with
and FeO are present in the capsule after the experiment, the Ir, resulting in the formation of needle-like aggregates,
no significant deviation in the imposed fO2 is expected. of which only few grains could be analysed (exp. 1583).
Our results indicate that the buffer capsule remains To improve the measurement statistics, the run time was
sealed, as Fe is incorporated in the Au, even at temper- increased by several hours. In IW-buffered experiments, this
atures near the nominal melting point of pure Au (i.e., led to more progressive reaction of Au with the buffer and
exp. 1581 at 1280 °C and 5 GPa, Table 1). This contrasts Fe gain in the surrounding sample material, respectively. To
with experiments at the higher fO2 of the Re–ReO2 buffer, solve the problem with the texture of the sensor platelets in
where Au–Re alloying causes capsule failure after < 12 h IW-buffered runs, we recommend using > 6 µm Ir powder
(Matjuschkin, unpub. data). However, we do observe Fe rather than increasing the run time.
diffusion through the buffer capsule (Fig.  6a). This is The EPMA measurements of the IrFe alloy reveal small
apparent in backscattered electron images as flame-like amounts of carbon, Au, and Ni (see supplementary infor-
Fe-rich zones emanating from the capsule into the adjacent mation). Gold is present in the sample in the form of small
olivine (Fig. 6b). Such a phenomenon needs to be carefully nuggets often attached to the sensor, which made it difficult
monitored to assure that the Fe enrichment does not reach to analyse. Such analyses were not taken into account when
the sample. This is an advantage of the olivine container determining the alloy composition. Ni contents vary from
in that it can readily incorporate this excess Fe (Fig. 6b). 0.3 to 2.5 wt% (see supplementary data).
An alternative is to use the Mo–MoO2 buffer, which lies Results of the sensor evaluation demonstrate a homoge-
only ~ 0.5 log units below IW (O’Neill 1986). In experi- neous fO2 distribution in the sample material, with ∆logfO2
ments conducted at the same pressure and temperature as values near IW + 0.5 (Table 1). If the small Ni impurity is
those with the IW buffer, we observed no migration of Mo considered together with Fe, the resulting alloy composition

13
1  Page 10 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

gives ∆logfO2 values that are shifted systematically to more


reducing conditions by < 0.2 log units. The IW buffer itself
is slightly more oxidised compared to Mo–MoO2 and, how-
ever, delivers similar results in experiments. This is likely
to be due to more effective buffering resulting from Fe infil-
tration from the buffer material. Thus, the IW buffer can
be used for faster achievement of fO2 equilibrium, whereas
the Mo–MO2 buffer can be employed, where a chemically
closed environment is required.

Discussion

Fluid speciation at fO2 ≈ IW buffer

Overall, run products and sensor readings demonstrate


achievement of strongly reduced conditions. Despite wide
differences in run times, temperatures and degree of Fe loss,
fluid compositions remained remarkably consistent. Metal
buffers helped to maintain the low fO2 environment; how-
ever, they experienced strong oxidation, by which up to 40%
of metal reacted to the metal -oxide. This is due to a high
demand for ­H2, which is > 7 times higher than more oxidised
Ni–NiO buffer conditions. Clearly, in the absence of metal
buffers, fluids would be unstable and would shift composi-
tionally towards the water maximum:
O2 + CH4 = C + 2H2 O, (6)
O2 + C2 H6 = 2C + 2H2 O + H2 , (7)
where the oxygen is produced by reaction 2 (also see Mat-
veev et al. 1997). Note that Fe loss to the metal is kinetically
slower compared to the fluid equilibration, Fe diffusion in
olivine or the metal–metal oxide buffer reaction (Matveev
et al. 1997).
Even though the fluid and fO2 controls appear to be suc-
cessful, the absence of water peaks in the Raman spectra is
somewhat unexpected. Our results agree with EOS calcula-
tion by Huizenga (2005) with ~ 3.7 mol% ­H2O, but contra-
dict predictions by Belonoshko and Saxena (1992), Matveev
et al. (1997), and Zhang and Duan (2010) suggesting up
to 32 mol% H ­ 2O at 5 GPa, 1280 °C, and IW conditions,
respectively. It also contrasts with fluid compositions given
Fig. 6  a Effect of Fe gain in the (filler) sample material placed adja- by Taylor and Foley (1989) who employed the WC–WO2–C
cent to the IW-buffer capsule (experiment 1581 5  GPa/1280  °C, assemblage to buffer fO2, and found measurable H ­ 2O by gas
24 h). Note that the Fe addition takes place along grain boundaries of chromatography. However, their experiments were reported
olivine and does not affect the orthopyroxene composition. b Effect
of Fe gain in IW-buffered experiments produces “flame-like” zones
to be actively undergoing oxidation during the experiment
of Fe-enriched olivine in the olivine container in experiment 1571. by ­H2 loss through the A­ g50Pd50 outer capsule. In addition,
c Reaction of Mo–MoO2 buffer in experiment at 1280  °C, 5 GPa the calculated fO2 of this buffer assemblage relative to IW
(experiment 1587). Note the absence of reaction between the buffer is based in part upon a fluid speciation model not unlike
and Au capsule. The presence of pore space indicates that H ­ 2O was
present inside the buffer capsule during the experiment
those mentioned above, adding significant uncertainty to the
actual redox conditions of these experiments (i.e., they could
easily have been somewhat more oxidising than thought).

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 11 of 14  1

Certainly, 10–100 s of ppm of water can be accommodated models are not accurate at more oxidising conditions. In fact,
as OH in olivine (Mosenfelder et al. 2006) and orthopy- it is likely that the models reliably reproduce fluid specia-
roxene (Rauch and Keppler 2002) but not in the volumes tion for fO2 conditions at and above the “water maximum”,
predicted by GFluid. In any case, high OH contents in the where compositions are dominated by ­H2O and C ­ O2 (e.g.,
silicate phases necessitate a correspondingly high H ­ 2O activ- Tiraboschi et al. 2016). However, the EOS and mixing prop-
ity in the coexisting fluid phase, which we do not observe. erties of C­ H4 are not as well constrained, making prediction
Finally, “unidentified” water might be present in small of fluid speciation under reducing conditions problematic.
amounts in fluid inclusions (< 5 mol%) and be simply invis-
ible to Raman spectroscopy (Lamadrid et al. 2014). In an Implications for the Earth’s mantle
attempt to increase the ­H2O concentration in the fluid, an
Mo-MoO2-buffered test experiment was performed during The low solubility of ­CH4 in minerals means that reduced
which it initially equilibrated at 7 GPa and 1250 °C and was ­CH4-bearing COH fluids have the potential of being stable
then subjected to a stepwise cooling down to 800 °C over at redox conditions observed for the deep portions of the
2 h. Even over this short time period, the fluid apparently upper mantle (Frost and McCammon 2008; Woodland and
re-equilibrated within the olivine-hosted inclusions, produc- Koch 2003) and are thus capable of transporting significant
ing measureable amounts of H ­ 2O along with C
­ H4 that were amounts of carbon. If at conditions of ∆logfO2 ≈ IW + 0.5
detectable in Raman spectra (Fig. 7). This result not only (at 1280 °C and 5 GPa), the fluid phase is essentially H ­ 2O
demonstrates that ­H2O can be readily detected when it is free (Table 1) as our results indicate, then the change in fluid
present, but it also serves as additional proof for the rapid composition during oxidation up to the “water maximum”
kinetics of fluid equilibration and the high sensitivity of fluid will be even more pronounced than predicted by available
speciation to the P, T, and fO2 conditions of an experiment. speciation models like GFluids. For example, at 1280 °C
Calculations with the GFluid model (Zhang and Duan and 5 GPa, the ­H2O/(CH4 + H2O) will shift from nearly 0
2009, 2010) and (Belonoshko and Saxena 1992) are in a to ~ 0.95 as the ∆logfO2 changes only 2.1 log units from
good agreement with high-pressure experiments by Mat- IW + 0.5 to IW + 2.6 (water max.), assuming that the GFluid
veev et al. (1997), who described that they had problems model is accurate at least for fO2 values down to the “water
of ­H2 loss, resulting in a shift in fluid composition towards maximum”. This means that more ­H2O will be produced per
the water maximum (more oxidised conditions). Hence, it is incremental increase in fO2 than previously thought (e.g.,
likely that our lower water contents are due to the improved Huizenga et al. 2012; Luth and Stachel 2014). In the context
capsule technique and represent more realistic values com- of redox melting, where H ­ 2O released during oxidation of a
pared to the earlier studies. As pointed out by Matveev et al. reduced fluid drives partial melting, our results imply that a
(1997) and demonstrated by Ardia et al. (2013), theoretical smaller degree of oxidation is necessary to produce a given
EOS calculations tend to overestimate water contents and degree of melting, since the ­H2O activity should increase
underestimate the amount of ­CH4 in fluids at high tempera- more rapidly with changing fO2. This process is driven
tures and pressures. This is not to say that such speciation by reaction (6) with the constraint that the surrounding

Fig. 7  CH4- and H­ 2O-bearing 7 GPa, 6h at 1250°C


fluid inclusions produced in an arbitary
+ 2h cooling to 800°C
experiment at 7 GPa, where the units
temperature was progressively
lowered during the experiment
from 1250 to 800 °C, see text
2917 CH4 (L)
5µm

10µm 2722 H2O


C

host olivine

1000 2000 3000 4000 cm-1

13
1  Page 12 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

peridotite is able to provide the necessary amount of ­O2, at the fO2 levels relevant for the mantle environment (i.e.,
which may not be the case (Luth and Stachel 2014). The Frost and McCammon 2008; Woodland and Koch 2003).
extent to which reaction (6) can proceed also has a direct
effect on the amount of graphite or diamond that may form. Acknowledgements  The Deutsche Forschungsgemeinschaft is grate-
fully acknowledged for funding the project WO652/26-1. This work
Alternatively, the oxidation of ­CH4 via the equilibrium: has benefited from discussions with Daniel J. Frost and Sonja Aulbach.
CH4 = C + 2H2 , (8)
can also contribute the crystallisation of graphite or diamond
without requiring a source of ­O2. In this case, the reaction
would shift from left to right through the preferential loss of References
­H2 from the immediate surroundings and no partial melting
Akella J, Kennedy GC (1971) Melting of gold, silver and copper—pro-
would necessarily be induced. posal for a new high-pressure calibration scale. J Geophys Res
The fluids in our experiments also contain unquantified 76(20):4969. https​://doi.org/10.1029/JB076​i020p​04969​
amounts of ­C2H6 and ­H2 (Fig. 5). The presence of ­H2 is Andersen DJ, Lindsley DH, Davidson PM (1993) QUILF: a pascal
indirectly indicated by detectable amounts of H ­ 2 having program to assess equilibria among Fe–Mg–Mn–Ti oxides, pyrox-
enes, olivine, and quartz. Comput Geosci 19(9):1333–1350
been incorporated into the crystal structure of the coexist- Arculus RJ (1985) Oxidation status of the mantle: past and present.
ing olivine (Fig. 5d). Under strongly reducing conditions, the Annu Rev Earth Planet Sci 13:75–95. https​://doi.org/10.1146/
incorporation of ­H2 in olivine and orthopyroxene provides annur​ev.ea.13.05018​5.00045​1
a local sink for ­H2. This stored ­H2 also has the potential to Ardia P, Hirschmann MM, Withers AC, Stanley BD (2013) Solubility
of ­CH4 in a synthetic basaltic melt, with applications to atmos-
produce ­H2O via reaction (1) during subsequent interactions phere-magma ocean-core partitioning of volatiles and to the evo-
with relatively oxidised mantle domains, further enhancing lution of the Martian atmosphere. Geochimica Et Cosmochimica
the redox-melting process. On the other hand, oxidation Acta 114:52–71. https​://doi.org/10.1016/j.gca.2013.03.028
would also cause an increase in OH defects in olivine and Ballhaus C, Berry RF, Green DH (1991) High pressure experimental
calibration of the olivine-orthopyroxene-spinel oxygen geoba-
other mantle phases (Yang et al. 2016), potentially compen- rometer: implications for the oxidation state of the upper mantle.
sating for this effect. Contrib Mineral Petrol 107:27–40
Barr JA, Grove TL (2010) AuPdFe ternary solution model and appli-
cations to understanding the fO(2) of hydrous, high-pressure
experiments. Contrib Mineral Petrol 160(5):631–643. https​://doi.
org/10.1007/s0041​0-010-0497-z
Conclusions Belonoshko AB, Saxena SK (1992) A unified equation of state for
fluids of C–H–O–N–S–Ar composition and their mixtures up to
A new triple-capsule design for high-pressure experiments very high temperatures and pressures. Geochimica Et Cosmo-
chimica Acta 56(10):3611–3626. https​://doi.org/10.1016/0016-
was tested. We have demonstrated its ability to attain and 7037(92)90157​-e
maintain the hydrogen fugacity (fH2) at levels corresponding Blundy J, Cashman K (2008) Petrologic reconstruction of magmatic
to Fe–FeO and Mo–MoO2 buffer conditions. The capsule system variables and processes. Miner Incl Volcan Process
design is suitable for experimental investigation of deep 69:179–239. https​://doi.org/10.2138/rmg.2008.69.6
Blundy J, Cashman KV, Rust A, Witham F (2010) A case for ­CO2-rich
mantle fluids, but has a temperature limit corresponding to arc magmas. Earth Planet Sci Lett 290(3–4):289–301. https​://doi.
the melting point of gold at the pressure and temperature of org/10.1016/j.epsl.2009.12.013
interest. An olivine container was used to prevent the Fe loss Boldyrev VV (2002a) Thermal decomposition of silver oxalate. Ther-
from the sample and as a trap for fluid inclusions. A C ­ aF2 mochim Acta 388(1–2):63–90. https​://doi.org/10.1016/s0040​
-6031(02)00044​-8
pressure-transmitting medium was used to minimise H ­ 2 loss Boldyrev (2002b) Thermal decomposition of silver oxylate. Ther-
from the capsule assembly and to promote faster fH2 equilib- mochim Acta 388:63–90. https ​ : //doi.org/10.1016/S0040​
rium in the assembly. To avoid contamination of the sample -6031(02)00044​-8
materials by the metal–metal oxide oxygen buffer, Fe or Mo Borisov A, Palme H (2000) Solubilities of noble metals in Fe-contain-
ing silicate melts as derived from experiments in Fe-free systems.
was packed into a gold capsule. Formation of an Au–Fe alloy Am Mineral 85(11–12):1665–1673
led to Fe gain; however, this can be minimised by employing Borisov A, Behrens H, Holtz F (2015) Effects of melt composition of
shorter run times of ~ 20 h. This is sufficient to reach fO2 ­Fe3+/Fe2+ in silicate melts: a step to model ferric/ferrous ratio in
equilibrium, provided that the mass of the sample is smaller multicomponent systems. Contrib Mineral Petrol 169:24. https​://
doi.org/10.1007/s0041​0-015-1119-6
than the mass of the buffer. The experimental fluid trapped Brey GP, Weber R, Nickel KG (1990) Calibration of a belt appara-
in inclusions in an olivine container at fO2 ≈ IW is domi- tus to 1800 °C and 6 GPa. J Geophys Res Solid Earth Planets
nantly ­CH4 and contains no detectable ­H2O. This confirms 95(B10):15603–15610. https:​ //doi.org/10.1029/JB095i​ B10p1​ 5603​
the success of our design being able to achieve and main- Cahn RW (1991) Binary alloy phase diagrams–second edition. T.
B. Massalski, Editor‐in‐Chief; H. Okamoto, P. R. Subrama-
tain reducing conditions in the presence of a mantle silicate nian, L. Kacprzak, Editors. ASM international, Materials Park,
assemblage. We suggest that methane-rich fluids can exist Ohio, USA. December 1990. xxii, 3589 pp., 3 vol., hard‐back.

13
Contributions to Mineralogy and Petrology (2019) 174:1 Page 13 of 14  1

$995.00 the set. Adv Mater 3:628–629. https​://doi.org/10.1002/ Lazarov M, Woodland AB, Brey GP (2009) Thermal state and redox
adma.19910​03151​2 conditions of the Kaapvaal mantle: a study of xenolith from the
Chou IM (1986) Permeability of precious metals to hydrogen Finsch mine, South Africa. Lithos 112S:913–923
at 2 kb total pressure and elevated temperatures. Am J Sci Luth RW, Stachel T (2014) The buffering capacity of lithospheric
286(8):638–658 mantle: imlications for diamond formation. Contrib Mineral
Christensen KO, Chen D, Lodeng R, Holmen A (2006) Effect of sup- Petrol 168:1083. https​://doi.org/10.1007/s0041​0-014-1083-6
ports and Ni crystal size on carbon formation and sintering dur- Massalski TB, Murray JL, Bennett LH, Baker H (1986) Binary alloy
ing steam methane reforming. Appl Catal A 314 (2006):9–22. phase diagrams. American Society for Metals, Metals Park
https​://doi.org/10.1016/j.apcat​a.2006.07.028 Matjuschkin V, Brooker RA, Tattitch B, Blundy JD, Stamper CC
Concha BE, Bartholomew GL, Bartholomew CH (1984) CO hydro- (2015) Control and monitoring of oxygen fugacity in piston
genation on supported molybdenum catalysts: effects of support cylinder experiments. Contrib Mineral Petrol. https​: //doi.
on specific activities of reduced and sulfided catalysts. J Catal org/10.1007/s0041​0-015-1105-z
89(2):536–541. https​://doi.org/10.1016/0021-9517(84)90332​-4 Matjuschkin V, Blundy JD, Brooker RA (2016) The effect of pres-
Eugster HP (1957) Heterogeneous reactions involving oxidation sure on sulphur speciation in mid-to deep-crustal arc magmas
and reduction at high pressures and temperatures. J Chem Phys and implications for the formation of porphyry copper depos-
26(6):1760–1761. https​://doi.org/10.1063/1.17436​26 its. Contrib Mineral Petrol. https​: //doi.org/10.1007/s0041​
Evans KA, Tomkins AG (2011) The relationship between subduction 0-016-1274-4
zone redox budget and arc magma fertility. Earth Planet Sci Lett Matveev S, Ballhaus C, Fricke K, Truckenbrodt J, Ziegenbein D
308(3–4):401–409. https​://doi.org/10.1016/j.epsl.2011.06.009 (1997) Volatiles in the Earth’s mantle.1. Synthesis of CHO
Evans WJ, Lipp MJ, Yoo C-S, Cynn H, Herberg JL, Maxwell RS, fluids at 1273 K and 2.4 GPa. Geochimica Et Cosmochim-
Nicol MF (2006) Pressure-induced polymerization of carbon ica Acta 61(15):3081–3088. https​: //doi.org/10.1016/s0016​
monoxide: disproportionation and synthesis of an energetic lac- -7037(97)00142​-7
tonic polymer. Chem Mater 18:2520–2531 Matzen AK, Baker MB, Beckett JR, Stolper EM (2011) Fe-Mg parti-
Freda C, Baker DR, Ottolini L (2001) Reduction of water loss from tioning between olivine and high-magnesian melts and the nature
gold-palladium capsules during piston-cylinder experiments by of Hawaiian parental liquids. J Petrol 52(7–8):1243–1263. https​
use of pyrophyllite powder. Am Mineral 86(3):234–237 ://doi.org/10.1093/petro​logy/egq08​9
French BM (1966) Some geological implications of equilibrium Melekhova E, Annen C, Blundy J (2013) Compositional gaps in igne-
between graphite and a C–H–O gas phase at high temperatures ous rock suites controlled by magma system heat and water con-
and pressures. Rev Geophys 4(2):223. https​://doi.org/10.1029/ tent. Nat Geosci 6(5):385–390. https​://doi.org/10.1038/ngeo1​781
RG004​i002p​00223​ Moore G, Roggensack K, Klonowski S (2008) A low-pressure-high-
Frost DJ, McCammon (2008) The redox state of Earth’s mantle. temperature technique for the piston-cylinder. Am Mineral
Annu Rev Earth Planet Sci 36:389–420 93(1):48–52. https​://doi.org/10.2138/am.2008.2618
Hanley L, Xu Z, Yates JT (1991) Methane activation on Ni(111) at Mosenfelder JL, Deligne NI, Asimow PD, Rossman GR (2006) Hydro-
high pressures. Surf Sci Lett 248:L265–L273 gen incorporation in olivine from 2 to 12 GPa. Am Mineral 91(2–
Holloway JR, Burnham CW, Millhollen GL (1968) Generation of 3):285–294. https​://doi.org/10.2138/am.2006.1943
­H2O–CO2 mixtures for use in hydrothermal experimentation. Mutch EJF, Blundy JD, Tattitch BC, Cooper FJ, Brooker RA (2016)
J Geophys Res 73:20 An experimental study of amphibole stability in low-pressure
Huizenga JM (2005) COH, an Excel spreadsheet for composi- granitic magmas and a revised Al-in-hornblende geobarometer.
tion calculations in the C–O–H fluid system. Comput Geosci Contrib Mineral Petrol 171(10) https​://doi.org/10.1007/s0041​
31(6):797–800. https​://doi.org/10.1016/j.cageo​.2005.03.003 0-016-1298-9
Huizenga JM, Crossingham A, Vijoen F (2012) Diamond precipita- O’Neill HStC (1986) Mo–MoO2 (MOM) oxygen buffer and the free
tion from ascending reduced fluids in the Kaapvaal lithosphere: energy of formation of ­MoO2. Am Mineral (71):1007–1010
thermodynamic constraints. CR Geosci 344:67–76 Pownceby MI, O’Neill HSC (1994a) Thermodynamic data from redox
Jakobsson S (2012) Oxygen fugacity control in piston-cylinder reactions at high-temperatures. III. Activity-composition rela-
experiments. Contrib Mineral Petrol 164(3) https ​ : //doi. tions in Ni–Pd alloys from EMF-measurements at 850–1250 K,
org/10.1007/s0041​0-012-0743-7 and calibration of the NiO–NiPd assemblage as a redox sensor.
Jakobsson S, Blundy J, Moore G (2014a) Oxygen fugacity control in Contrib Mineral Petrol 116(3):327–339. https​://doi.org/10.1007/
piston-cylinder experiments: a re-evaluation. Contrib Mineral bf003​06501​
Petrol 167(6) https​://doi.org/10.1007/s0041​0-014-1007-5 Pownceby MI, O’Neill HSC (1994b) Thermodynamic data from redox
Jakobsson S, Blundy J, Moore G (2014b) Oxygen fugacity control in reactions at high-temperatures. IV. Calibration of the Re–ReO2
piston-cylinder experiments: a re-evaluation. Contrib Mineral oxygen buffer from Emf and NiO–Ni–Pd redox sensor meas-
Petrol 167(1007) https​://doi.org/10.1007/s0041​0-014-1007-5 urements. Contrib Mineral Petrol 118(2):130–137. https​://doi.
Jamieson HE, Roeder PL, Grant AH (1992) Olivine–pyroxene–Pt– org/10.1007/bf010​52864​
FeL alloy as an oxygen geobarometer. J Geol 100(1):138–145 Pownceby MI, O’Neill HSC (1995) Thermodynamic data from redox
Kratzer P, Hammer B, Nørskov JK (1996) A theoretical study of C ­ H4 reactions at high-temperatures. V. Thermodynamic properties of
dissociation on pure and gold-alloyed Ni(111) surfaces. J Chem NiO-MnO solid solutions from EMF-measurements. Contrib Min-
Phys 105(13):5595–5604 eral Petrol 119(4):409–421
Kress VC, Carmichael ISE (1991) The compressibility of silicate Pownceby MI, O’Neill HSC (2000) Thermodynamic data from redox
liquids containing ­Fe2O3 and the effect of composition, tem- reactions at high temperatures. VI. Thermodynamic properties of
perature, oxygen fugacity and pressure on their redox states. CoO–MnO solid solutions from emf measurements. Contrib Min-
Contrib Mineral Petrol 108:82–92 eral Petrol 140(1):28–39. https​://doi.org/10.1007/s0041​00000​162
Lamadrid HM, Lamb WM, Santosh M, Bodnar RJ (2014) Raman Ratajeski K, Sisson TW (1999) Loss of iron to gold capsules in rock-
spectroscopic characterization of ­H2O in C ­ O2-rich fluid inclu- melting experiments. Am Miner 84(10):1521–1527
sions in granulite facies metamorphic rocks. Gondwana Res Rauch M, Keppler H (2002) Water solubility in orthopyroxene. Contrib
26:301–310 Mineral Petrol 143:525–536

13
1  Page 14 of 14 Contributions to Mineralogy and Petrology (2019) 174:1

Rosenbaum JM, Slagel MM (1995) C–O–H Speciation in piston-cyl- Truckenbrodt J, Ziegenbein D, Johannes W (1997) Redox condi-
inder experiments. Am Miner 80(1–2):109–114 tions in piston-cylinder apparatus: the different behavior of
Rostrup-Nielsen JR (1993) Production of synthesis gas. Catal Today boron nitride and unfired pyrophyllite assemblies. Am Mineral
18(4):305–324. https​://doi.org/10.1016/0920-5861(93)80059​-a 82(3–4):337–344
Serra E, Bini AC, Cosoli G, Pilloni L (2005) Hydrogen permeation Uenver-Thiele L, Woodland AB, Seitz HM, Downes H, Altherr R
measurements on alumina. J Am Ceram Soc 88(1):15–18. https​ (2017) Metasomatic processes revealed by trace element and
://doi.org/10.1111/j.1551-2916.2004.00003​.x redox signatures of the lithospheric mantle beneath the Massif
Sisson TW, Ratajeski K, Hankins WB, Glazner AF (2005) Volumi- Central, France. J Petrol 58(3):395–422. https​://doi.org/10.1093/
nous granitic magmas from common basaltic sources. Contrib petro​logy/egx02​0
Mineral Petrol 148(6):635–661. https​://doi.org/10.1007/s0041​ Vannice MA (1977) Catalytic synthesis of hydrocarbons from ­H2-CO
0-004-0632-9 mixtures over group VIII metals. V. Catalytic behaviour of
Sokol AG, Palyanova GA, Palyanov YN, Tomilenko AA, Mele- silica supported metals. J Catal 50(2):228–236. https​: //doi.
nevsky VN (2009) Fluid regime and diamond formation in the org/10.1016/0021-9517(77)90031​-8
reduced mantle: experimental constraints. Geochimica Et Cos- Woermann E, Rosenhauer M (1985) Fluid phases and the redox state
mochimica Acta 73(19):5820–5834. https​://doi.org/10.1016/j. of the Earth’s mantle: explorations based on experimental, phase-
gca.2009.06.010 theoretical and petrological data. Fortschr Mineral 63(2):263–349
Sokol AG, Tomilenko AA, Bul’bak TA, Palyanova GA, Sokol IA, Wood BJ (1990) An experimental test of the spinel peridotite oxygen
Palyanov YN (2017) Carbon and nitrogen speciation in N-poor barometer. J Geophys Res 95(B10):15845–15851
C–O–H–N fluids at 6.3 GPa and 1100–1400 degrees C. Sci Rep. Woodland AB, Koch M (2003) Variation in oxygen fugacity with depth
https​://doi.org/10.1038/s4159​8-017-00679​-7 in the upper mantle beneath the Kaapvaal Craton, southern Africa.
Solymosi F, Erdohelyi A (1980) Hydrogenation of ­CO2 to ­CH4 over Earth Planet Sci Lett 214:295–310
alumina-supported noble metalsHydrogenation of ­CO2 to ­CH4 Woodland AB, O’Neill HS (1997) Thermodynamic data for Fe-bearing
over alumina-supported noble metals. J Mol Catal 8(4):471–474. phases obtained using noble metal alloys as redox sensors. Geo-
https​://doi.org/10.1016/0304-5102(80)80086​-1 chimica Et Cosmochimica Acta 61(20):4359–4366. https​://doi.
Stamper CC, Melekhova E, Blundy JD, Arculus RJ, Humphreys org/10.1016/s0016​-7037(97)00247​-0
MCS, Brooker RA (2014) Oxidised phase relations of a primi- Yang X, Keppler H, Li Y (2016) Molecular hydrogen in mantle miener-
tive basalt from Grenada, Lesser Antilles. Contrib Mineral Petrol als. Geochem Perspect 2:160–168. https​://doi.org/10.7185/geoch​
167(1):1–20 emlet​.1616
Taylor WR, Foley SF (1989) Improved oxygen-buffering techniques Zellmer GF, Edmonds M, Straub SM (2015) Volatiles in subduction
for C–O–H fluid-saturated experiments at high-pressure. J Geo- zone magmatism. Role of volatiles in the genesis. Evol Erupt Arc
phys Res Solid Earth Planets 94(B4):4146–4158. https​://doi. Magmas 410:1–17. https​://doi.org/10.1144/sp410​.13
org/10.1029/JB094​iB04p​04146​ Zhang C, Duan Z (2009) A model for C-O-H fluid in the Earth’s
Taylor WR, Green DH (1987) The petrogenetic role of methane: mantle. Geochim Cosmochim Acta 73:2089–2102. https​://doi.
effect on liquidus phase relations and the solubility mechanism org/10.1016/j.gca.2009.01.021
of reduced C–H volatiles. In: Mysen BO (ed) Magmatic processes: Zhang C, Duan ZH (2010) GFluid: an Excel spreadsheet for investigat-
physicochemical principles. Geochemical Society, Pennsylavania ing C–O–H fluid composition under high temperatures and pres-
State University, Philadelphia, pp 121–138 sures. Comput Geosci 36(4):569–572. https​://doi.org/10.1016/j.
Taylor JR, Wall VJ, Pownceby MI (1992) The calibration and appli- cageo​.2009.05.008
cation of accurate redox sensors. Am Mineral 77(3–4):284–295
Tiraboschi C, Tumiati S, Recchia S, Miozzi F, Poli S (2016) Quantita- Publisher’s Note Springer Nature remains neutral with regard to
tive analysis of COH fluids synthesized at HP–HT conditions: an jurisdictional claims in published maps and institutional affiliations.
optimized methodology to measure volatiles in experimental cap-
sules. Geofluids 16(5):841–855. https:​ //doi.org/10.1111/gfl.12191​
Truckenbrodt J, Johannes W (1999) ­H2O loss during piston-cylinder
experiments. Am Miner 84(9):1333–1335

13

You might also like