ArticlePDF Available

Abstract and Figures

Fourier Transform infrared (FTIR) absorption spectra of hydroxyl were measured on olivine phenocrysts from hydrous basaltic melts that originated in island-arc tectonic settings. The basaltic melts encom-pass a wide range of silica activities from orthopyroxene-saturated hypersthene-normative to nepheline-normative compositions. The intensities and wavenumber placement of hydroxyl absorption bands correlate with the degree of silica saturation of the parent melt from which the olivine crystallized. Olivines from silica-undersaturated nepheline-normative melts absorb IR radiation in the wavenumber range 3430–3590 cm-1 (Group 1). In contrast, olivines from orthopyroxene-saturated boninitic melts exhibit hydroxyl absorption bands in the wavenumber range 3285–3380 cm-1 (Group 2). Olivines crystallized at intermediate silica activities exhibit a combina-tion of the two groups of hydroxyl IR bands, where the proportion of Group 2 bands increases with increasing silica saturation of the parent melt. The positions of hydroxyl absorption peaks observed here for natural samples are consistent with previous measurements on experimentally annealed olivines. Thus protonation experiments can be employed to make spectroscopically dry olivine structures visible by IR, yielding information on the silica saturation of the parental magmas. Hydroxyl concentrations in the studied olivines were estim-ated to be 1–2 ppm, corresponding to an olivine–melt partition coefficient of �(1�.0 �+/- 0�.3) �x10�4.
Content may be subject to copyright.
FTIR Spectrum of Phenocryst Olivine as an
Indicator of Silica Saturation in Magmas
S. MATVEEV
1,2
*, M. PORTNYAGIN
3
, C. BALLHAUS
2
, R. BROOKER
4
AND C. A. GEIGER
5
1
DEPARTMENT OF EARTH AND ATMOSPHERIC SCIENCES, UNIVERSITY OF ALBERTA,
1–26 EARTH SCIENCES BUILDING, EDMONTON, ALTA., TG6 2E3, CANADA
2
INSTITUT FU
¨R MINERALOGIE, UNIVERSITA
¨TMU
¨NSTER, CORRENSSTR. 24, 48149 MU
¨NSTER, GERMANY
3
LEIBNIZ INSTITUTE FOR MARINE SCIENCES, DYNAMICS OF THE OCEAN FLOOR, WISCHHOFSTR. 1–3, 24148 KIEL,
GERMANY
4
DEPARTMENT OF EARTH SCIENCES, UNIVERSITY OF BRISTOL, QUEENS ROAD, BRISTOL BS8 1RJ, UK
5
INSTITUT FU
¨R GEOWISSENSCHAFTEN, UNIVERSITA
¨T KIEL, LUDEWIG-MEYN-STRAßE 10, 24118 KIEL, GERMANY
RECEIVED SEPTEMBER 15, 2003; ACCEPTED OCTOBER 7, 2004
ADVANCE ACCESS PUBLICATION DECEMBER 3, 2004
Fourier Transform infrared (FTIR) absorption spectra of hydroxyl
were measured on olivine phenocrysts from hydrous basaltic melts that
originated in island-arc tectonic settings. The basaltic melts encom-
pass a wide range of silica activities from orthopyroxene-saturated
hypersthene-normative to nepheline-normative compositions. The
intensities and wavenumber placement of hydroxyl absorption bands
correlate with the degree of silica saturation of the parent melt from
which the olivine crystallized. Olivines from silica-undersaturated
nepheline-normative melts absorb IR radiation in the wavenumber
range 3430–3590 cm
1
(Group 1). In contrast, olivines from
orthopyroxene-saturated boninitic melts exhibit hydroxyl absorption
bands in the wavenumber range 3285–3380 cm
1
(Group 2).
Olivines crystallized at intermediate silica activities exhibit a combina-
tion of the two groups of hydroxyl IR bands, where the proportion of
Group 2 bands increases with increasing silica saturation of the
parent melt. The positions of hydroxyl absorption peaks observed
here for natural samples are consistent with previous measurements on
experimentally annealed olivines. Thus protonation experiments can
be employed to make spectroscopically dry olivine structures visible by
IR, yielding information on the silica saturation of the parental
magmas. Hydroxyl concentrations in the studied olivines were estim-
ated to be 1–2 ppm, corresponding to an olivine–melt partition
coefficient of (1003) 10
4
.
KEY WORDS: nominally anhydrous minerals; olivine; water; mantle; silica
activity; melt inclusions
INTRODUCTION
Following the recognition that nominally anhydrous
rock-forming silicates can incorporate small amounts of
hydrogen (Beran, 1970; Wilkins & Sabine, 1973), much
work has been undertaken to document the hydroxyl
concentrations and possible hydroxyl substitution
mechanisms in olivine using Fourier Transform infrared
(FTIR) spectroscopy (e.g. Beran & Putnis, 1983; Aines &
Rossman, 1984; Miller et al., 1987; Mackwell &
Kohlstedt, 1990; Bai & Kohlstedt, 1993; Kohlstedt et al.,
1996; Kohn, 1996; Libowitzky & Rossman, 1997; Ingrin
& Skogby, 2000; Matveev et al., 2001; Lemaire et al.,
2004), and also using computational methods ( Wright &
Catlow, 1994; Braithwaite et al., 2003). FTIR spectro-
scopic studies of natural olivine show that hydroxyl
stretching bands generally occur in two wavenumber
regions, one between 3430 and 3630 cm
1
and the other
between 3285 and 3380 cm
1
. These are, following the
work of Bai & Kohlstedt (1993), referred to as Group 1
and Group 2 hydroxyl bands, respectively. In addition,
synthetic pure forsterite and rare natural samples
show lower-frequency hydroxyl bands ranging down to
3100 cm
1
(Miller et al., 1987; Demouchy & Mackwell,
2003; Lemaire et al., 2004).
Matveev et al. (2001) have shown experimentally
that the IR absorption of dissolved hydroxyl relates to
the silica activity (a
SiO
2
) at which the natural olivine
*Corresponding author. Present address: Department of Earth and
Atmospheric Sciences, University of Alberta, 1–26 Earth Sciences
Building, Edmonton, Alta., TG6 2E3, Canada. Telephone: þþ1 780
492 3191. Fax: þþ1 780 492 2030. E-mail: smatveev@ualberta.ca
#The Author 2004. Published by Oxford University Press. All
rights reserved. For Permissions, please email: journals.permissions@
oupjournals.org
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 PAGES 603–614 2005 doi:10.1093/petrology/egh090
crystallized or finally equilibrated. When olivine is equilib-
rated with magnesiowuustite (i.e. at low a
SiO
2
), the hydroxyl
stretching bands occur between 3430 and 3630 cm
1
(Group 1). On the other hand, when olivine is equilibrated
with orthopyroxene (i.e. at high a
SiO
2
), the hydroxyl bands
occur at lower energies between 3285 and 3380 cm
1
(Group 2). Hydroxyl in olivine could be associated with
the presence of cation vacancies whose concentrations
are largely controlled by a
SiO
2
(Stocker & Smyth, 1978;
Nakamura & Schmalzried, 1983), whereby with decreas-
ing a
SiO
2
the concentration of silicon vacancies increases
and the concentration of metal vacancies decreases.
Under this premise, Matveev et al. (2001) assigned
Group 1 absorption bands to hydroxyl groups associated
with silicon vacancies (i.e. a hydrogarnet-type substitu-
tion) and Group 2 absorption bands to hydroxyl asso-
ciated with octahedral M-site (Fe, Mg) vacancies. Thus
the possibility of locating hydrogen in silicon vacancies
might stabilize these defects and promote their formation
during olivine crystallization at low a
SiO
2
, even though the
concentration of silicon vacancies in the anhydrous oli-
vine structure might be regarded as insignificant (e.g.
Nakamura & Schmalzried, 1983; Mackwell et al., 1988).
These experimental results are supported by recent
computer simulations undertaken for the pure forsterite
(Mg
2
SiO
4
) crystal structure (Braithwaite et al., 2003). The
simulations were made to determine the energetics asso-
ciated with various possible hydroxyl substitutional
mechanisms, including both cation vacancies and inter-
stitials (nominally unoccupied crystallographic positions).
Braithwaite et al. concluded that hydroxyl groups should
occur in either vacant octahedral or tetrahedral sites so as
to produce neutral defect complexes, which are energe-
tically favorable relative to other substitution mechan-
isms. The simulations indicate that hydroxyl stretching
bands occur in two wavenumber regions. Their calcu-
lated spectra show that in the case where Si is replaced by
four H
þ
atoms, four hydroxyl stretching bands occur at
relatively high wavenumbers, whereas in the case where
two H
þ
atoms substitute for one Mg, two hydroxyl bands
occur at lower energies. The calculations agree with the
experimental observations of Lamaire et al. (2004), which
showed a significant effect of a
SiO
2
on the FTIR spectrum
of hydroxyl-bearing synthetic forsterite. These results are
in line with the experimental observations of Matveev
et al. (2001), although quantitative comparisons of
OH stretching frequencies measured on pure forsterite
with data obtained for natural samples are not possible
because of the effects of Fe and trace elements, such as Ti
(Berry et al., 2004), on the defect structure of olivine.
In this paper, we test the postulate of Matveev et al.
(2001) that the hydroxyl substitution mechanism could
serve as an indicator of the silica activity at which the
crystal structure of natural olivine equilibrates. To do this,
we measured the FTIR absorption spectra of hydroxyl in
olivine phenocrysts from a range of primitive mafic vol-
canic rocks. From bulk-rock compositions as well as the
composition of olivine melt inclusions, the parental melts
were all water-bearing and cover a wide range in a
SiO
2
.
SAMPLES
The basaltic lavas from which olivines were separated are
listed in Table 1. The eight samples are volcanic rocks
from subduction-related tectonic settings and include
lavas from (1) Troodos, Cyprus, (2) Avacha Volcano,
Kamchatka, Russia, and (3) Mount Mahimba, New
Georgia Archipelago, Solomon Islands. All the samples
are primitive in the sense that the melts from which
olivine phenocrysts crystallized could be in Fe–Mg
exchange equilibrium with typical mantle olivine (Fo
88–92
).
In the most silica-rich samples from Cyprus, olivine is
joined by orthopyroxene, and in the samples from
Avacha volcano and the Solomon Islands clinopyroxene
is the early liquidus mineral along with olivine.
Compositions of basaltic lavas and homogenized oli-
vine melt inclusions (e.g. Danyushevsky et al., 2002) are
plotted in Fig. 1 on projections from the olivine (a) and
diopside (b) apices onto the base of the ‘basalt tetrahe-
dron’. The molar concentrations of normative olivine
(Ol), diopside (Di), quartz (Qtz), jadeite ( Jd), Ca-Tscher-
mak pyroxene (CaTs), and leucite (Lc) were calculated
following procedures given by Falloon & Green (1987)
(Tables 2 and 3). To quantify the silica saturation degree
of the primary melts we use projections of plotted composi-
tions on the ( Jd þCaTs þLc)(Qtz) join parameter-
ized as the Qtz/(Qtz þJd þCaTs þLc) molar ratio,
subsequently referred to as the Silica Saturation Index
(SSI) (Tables 2 and 3). The SSI is independent of changes
in normative olivine and diopside concentrations, and
thus remains largely unchanged during magma fractio-
nation or during reaction of melt with host olivine. Thus
the SSI of melt inclusions or basaltic lavas adequately
reflects the silica-saturation degree of the primary melt.
Another advantage of SSI is that unlike a
SiO
2
it can be
easily calculated for magmas where olivine is not coexist-
ing with orthopyroxene or magnesiowuustite (compare
Carmichael, 2002). A negative correlation between SSI
and TiO
2
concentration in both basaltic lavas and melt
inclusions is consistent with the origin of the Troodos
primary magmas from variably depleted sources as a
result of different degrees of melting (Cameron, 1985;
Sobolevet al., 1993;Falloon & Danyushevsky,2000) (Fig. 2).
Lavas from the Troodos ophiolite and
Mamonia complex, Cyprus
The Troodos ophiolite and disaggregated ophiolitic
blocks in the Mamonia complex in Cyprus are
tectonically exhumed fragments of Upper Cretaceous
604
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
oceanic-like crust interpreted to have been formed in a
supra-subduction zone (Pearce, 1975; Cameron, 1985;
Rautenschlein et al., 1985). They represent the ‘birth’
and ‘youth’ stages of island-arc volcanism (Shervais,
2001). During the initial stages of subduction the astheno-
spheric mantle sources produced high-magnesian
hypersthene- to quartz-normative basaltic melts including
island-arc tholeiites and refractory high-CaO boninites
(Cameron, 1985; Sobolev et al., 1993). The presence of
water and hydroxyl in pillow-rim glasses and in melt
inclusions in phenocrysts suggests involvement of water
during mantle melting (e.g. Rautenschlein et al., 1985;
Sobolev & Chaussidon, 1996). In the samples studied,
bulk-rock MgO ranges from 22 to 327 wt %, and the
Fig. 1. Compositions of basaltic lavas (&) and average compositions of melt inclusions (*) shown on projections onto the base of the normative
‘basalt tetrahedron’ (Falloon & Green, 1987) from olivine (left diagram) and diopside (right diagram) apices. The outlined field is the entire
compositional range of melt inclusions. Relative proportions of ( Jd þCaTs þLc) and (Qtz) components define the Silica Saturation Index (SSI)
of magmas. Silica-undersaturated magmas with SSI <06 crystallize olivine exhibiting only the Group 1 hydroxyl absorption bands. Olivine from
orthopyroxene-saturated magmas with SSI >075 exhibits only the Group 2 hydroxyl bands. Shaded sector corresponds to the SSI from 060 to
075 where olivines exhibit both Group 1 and Group 2 absorption bands.
Table 1: Sample description
Sample Location Geochemical type Phenocryst mineralogy Fo range (mol %) Reference
MAM-25 Mavrokolimbos Dam,
Mamonia complex, Cyprus
High-Ca boninite Ol þCrSp þ(Opx
*
)87
.091.5 Portnyagin et al. (1996)
TRD-64 Kalavasos village,
Limassol Forest complex, Cyprus
High-Ca boninite Ol þOpx þCrSp 89.091.5 Sobolev et al. (1993)
TRD-41 Margi village, Troodos ophiolite,
Cyprus
Transitional to
high-Ca boninite
Ol þCrSp 89.093.5 Sobolev et al. (1993)
TRD-39 Pedieous River valley,
Troodos ophiolite, Cyprus
Transitional to
high-Ca boninite
Ol þCpx þCrSp 88.392.5 Sobolev et al. (1993)
TRD-75 Margi village, Troodos ophiolite,
Cyprus
Transitional to
high-Ca boninite
Ol þCrSp 89.291.2 Sobolev et al. (1993)
TRD-150 Analiondas village,
Troodos ophiolite, Cyprus
Low-K tholeiite Ol þCpx þCrSp 81.791.7 Portnyagin et al. (1997)
AV-2 Avacha volcano, Kamchatka, Russia Calc-alkaline basalt Ol þCpx þPl þCrSp 85.291.0 Portnyagin et al. (2004)
SS-1 Mt. Mahimba, Kolo Caldera,
New Georgia Island, Solomon Islands
Picrite Ol þCpx þCrSp 9092 Rohrbach (2003)
*
Opx found as inclusions in olivine.
605
MATVEEV et al. FTIR SPECTRUM OF OLIVINE
lavas contain up to 50 vol. % of 05–3 cm diameter
euhedral olivine phenocrysts. The unusually MgO-rich
samples were shown by Sobolev et al. (1993) to be olivine
cumulates in an evolved melt matrix.
Previously reported compositions of rocks and homo-
genized melt inclusions from forsteritic olivine pheno-
crysts (>Fo
88
) are summarized in Tables 2 and 3. Water
contents are based on ion-probe analysis of homogenized
melt inclusions and may be as high as 25 wt % in the
most silica-rich, refractory boninitic melts. Further details
concerning the petrography of the samples and the melt
inclusion data have been given by Sobolev et al. (1993),
Portnyagin et al. (1996, 1997), Sobolev & Chaussidon
(1996) and Portnyagin (1998).
Lava from Avacha volcano
Basaltic lava AV-2, from Avacha volcano in the
Kamchatka arc, with 15 wt % MgO (Tables 2 and
3) contains about 35 vol. % of large (up to 2 cm) pheno-
crysts of magnesian olivine (Fo
91–80
) and clinopyroxene
(Mg-number 92–73). This represents one of the most
primitive basalts of the Quaternary volcanic series of
the Kamchatka arc. Based on the compositions of homo-
genized melt inclusions in olivine, the parental magmas
were strongly silica undersaturated under crustal condi-
tions, and were probably derived from a refractory
asthenospheric mantle source that was refertilized prior
to, or during, partial melting by a subduction-related
component rich in H
2
O, CO
2
, and incompatible
trace elements. In contrast to the parental melt composi-
tion inferred from melt inclusions, the host lava is
hypersthene-normative and was interpreted to be a
mush of olivine and clinopyroxene crystals in an evolved
andesitic matrix (Portnyagin et al., 2004).
Lava from the Solomon Islands
Sample SS-1 is an olivine clinopyroxene-phyric lava
with 24 wt % MgO, derived from Mt. Mahimba, Kolo
Caldera, New Georgia Island. The sample is a cumulate
derived from a highly oxidized olivine–hypersthene-
normative parental melt with 132 wt % MgO and
489 wt % SiO
2
. Rohrbach (2003) identified two
Table 2: Average compositions of rocks
Sample MAM-25 TRD-64 TRD-41 TRD-39 TRD-75 TRD-150 AV-2 SS-1
wt %
SiO
2
45.66 42.81 44.64 42.13 41.72 44.93 50.89 48.90
TiO
2
0.16 0.12 0.26 0.18 0.23 0.33 0.52 0.61
Al
2
O
3
7.26 4.93 6.97 4.95 5.16 10.12 9.68 11.06
FeO 9.19 8.62 8.17 9.05 9.15 8.21 8.14 11.40
MnO 0.20
.15 0.15 0.17 0.16 0.16 0.16
MgO 26.48 32.52 28.34 31.44 32.73 22.00 16.10 13.2
CaO 5.15 3.68 5.01 3.35 2.61 8.09 11.56 9.75
Na
2
O0
.23 0.18 0.63 0.25 0.11 0.46 1.57 1.80
K
2
O0
.05 0.05 0.25 0.08 0.07 0.06 0.36 0.83
P
2
O
5
0.04 0.02 0.03 0.03 0.04 0.04 0.09
Cr
2
O
3
0.23 0.29 0.35 0.25 0.25 0.14
LOI 4.00 5.79 4.33 7.94 7.58 5.79 0.75
Total 98.65 99.14 99.11 99.80 99.79 100.30 99.82 97.35
Normative components
Jd 0.017 0.013 0.046 0.019 0.008 0.034 0.128 0.128
Lc 0.002 0.002 0.012 0.004 0.004 0.003 0.017 0.039
CTTs 0.005 0.003 0.007 0.005 0.007 0.009 0.015 0.017
CaTs 0.149 0.102 0.120 0.099 0.107 0.200 0.138 0.139
Di 0.057 0.048 0.077 0.038 0.000 0.122 0.297 0.228
Ol 0.582 0.707 0.600 0.707 0.742 0.466 0.239 0.317
Qtz 0.188 0.125 0.137 0.128 0.133 0.167 0.166 0.132
SSI 0.771 0.762 0.698 0.760 0.770 0.679 0.638 0.563
Data source references are given in Table 1. Normative components (molar fractions): Jd ¼NaAlSi
2
O
6
,Lc¼KAlSi
2
O
6
,
CTTs ¼CaTiAl
2
O
6
, CaTs ¼CaAl
2
SiO
6
,Di¼Ca(Mg,Fe,Mn)Si
2
O
6
,Ol¼(Mg,Fe,Mn)
3
Si
15
O
6
, Qtz ¼Si
3
O
6
. SSI (Silica
Saturation Index) ¼Qtz/(Qtz þCaTs þLc þJd) (see text for details). LOI, loss on ignition.
606
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
generations of primitive olivine phenocrysts (Fo
90–92
),
one with 01 and the other with >03 wt % CaO, of
which only the high-CaO generation is considered to
have been in equilibrium with the CaO content of
the parent melt. The low-CaO olivines are probably
xenocrysts from the lithospheric upper mantle. Con-
sequently, FTIR spectra were collected from only the
high-CaO olivines.
PROTONATION EXPERIMENTS
To make the defect structure of anhydrous olivine
visible by IR spectroscopy, olivine can be protonated
experimentally (Bai & Kohlstedt, 1993). Unlike equilibra-
tion experiments in which the olivine defect structure is
largely controlled by buffered a
SiO
2
(e.g. Bai & Kohlstedt,
1993; Kohlstedt et al., 1996; Matveev et al., 2001), the
protonation experiments are designed to keep the origi-
nal olivine defect structure intact. Therefore, pressure–
temperature conditions and run durations must be such
that the relaxation time of cation point defect equilibra-
tion significantly exceeds the run time of an experiment.
To optimize experimental conditions, we have used the
diffusion coefficients for hydrogen and metal vacancies
reported by Kohlstedt & Mackwell (1998) and references
therein. We also assumed that the diffusion rate of silicon
vacancies does not exceed that of metal vacancies
Table 3: Representative compositions of melt inclusions in olivine
MAM-25 TRD-64 TRD-39 TRD-41 TRD-75 TRD-150 AV-2 SS-1
n: 9 14 15 60 28 103 15 1
wt %
SiO
2
53.12 0.62 52.98 0.58 52.49 0.50 52.30 0.52 52.00 0.46 51.42 0.17 49.26 0.87 n.d.
TiO
2
0.19 0.02 0.20 0.01 0.38 0.04 0.38 0.02 0.48 0.03 0.53 0.02 0.99 0.16 n.d.
Al
2
O
3
12.12 0.28 12.24 0.46 11.53 0.42 12.80 0.39 13.27 0.37 15.58 0.17 14.42 0.73 n.d.
FeO 5.80 0.22 6.18 0.28 7.34 0.32 7.04 0.28 6.73 0.23 6.20 0.11 6.20 0.20 n.d.
MnO 0.13 0.02 0.14 0.02 0.16 0.01 0.14 0.01 0.13 0.01 0.14 0.01 0.11 0.01 n.d.
MgO 14.66 0.80 15.17 0.79 14.13 1.08 14.72 1.31 13.21 0.91 9.97 0.30 11.59 1.09 n.d.
CaO 11.09 0.63 11.35 0.36 10.31 0.48 9.91 0.44 10.37 0.27 12.83 0.17 13.81 0.94 n.d.
Na
2
O0
.55 0.15 0.42 0.03 1.30 0.05 1.50 0.10 1.77 0.10 1.40 0.04 2.65 0.17 n.d.
K
2
O0
.09 0.03 0.08 0.01 0.16 0.02 0.16 0.02 0.19 0.04 0.11 0.01 0.68 0.07 n.d.
H
2
O1
.73 0.03 1.56 0.17 1.85 0.19 1.71 0.09 1.63 0.06 1.36 0.05 <0.1 n.d.
Total 99.48 0.47 100.32 0.40 99.65 0.59 100.67 0.33 99.79 0.29 99.53 0.16 99.70 0.27
Normative components
Jd 0.038 0.010 0.029 0.002 0.104 0.004 0.091 0.007 0.124 0.006 0.098 0.003 0.209 0.015
Lc 0.004 0.001 0.003 0.001 0.007 0.001 0.007 0.001 0.009 0.002 0.005 0.001 0.030 0.002
CTTs 0.005 0.001 0.005 0.001 0.010 0.001 0.010 0.001 0.013 0.001 0.014 0.001 0.025 0.002
CaTs 0.230 0.004 0.235 0.008 0.204 0.009 0.186 0.007 0.202 0.006 0.265 0.003 0.183 0.010
Di 0.191 0.025 0.192 0.012 0.165 0.019 0.204 0.013 0.185 0.008 0.216 0.005 0.289 0.029
Ol 0.257 0.018 0.267 0.019 0.278 0.023 0.262 0.030 0.244 0.019 0.170 0.004 0.169 0.013
Qtz 0.275 0.012 0.267 0.009 0.232 0.011 0.239 0.010 0.224 0.008 0.232 0.003 0.096 0.019
SSI 0.751 0.007 0.749 0.005 0.687 0.006 0.715 0.006 0.667 0.006 0.654 0.003 0.371 0.048
Selected trace element compositions (ppm)
Cr 2400 3080 2700 2500 2860 1450 1240
Ni 581 1435 1243 1005 931 581 360
Co 76 103 105 93 108 76 48
Zn 59 64 61 66 66 59
Zr14 4251416 7 35
Errors calculated at 95% confidence level for ninclusions analysed. Data source references are given in Table 1. Normative
components (molar fractions): Jd ¼NaAlSi
2
O
6
,Lc¼KAlSi
2
O
6
, CTTs ¼CaTiAl
2
O
6
, CaTs ¼CaAl
2
SiO
6
,Di¼
Ca(Mg,Fe,Mn)Si
2
O
6
,Ol¼(Mg,Fe,Mn)
3
Si
1.5
O
6
, Qtz ¼Si
3
O
6
. SSI (Silica Saturation Index) ¼Qtz/(Qtz þCaTs þLc þJd)
(see text for details). n.d., not determined.
607
MATVEEV et al. FTIR SPECTRUM OF OLIVINE
(Mackwell et al., 1988; Matveev et al., 2001). Experiments
were performed in a piston-cylinder apparatus and held
at 1000C and 2 GPa for 4 h. At these run conditions, we
calculate that hydrogen may penetrate and protonate a
05–1 mm diameter olivine grain (the typical grain size of
our olivine separates), with the defect structure remaining
intact.
For the protonation experiments we have chosen
olivines from Solomon Islands basalt (SS-1) in which
the primary hydroxyl content was found to be below
the detection limit of the FTIR analysis. To ensure
that the olivine defect structure did not change in the
course of protonation we performed two a
SiO
2
-buffered
experiments: in one, olivine crystals were embedded in
periclase powder; in the other, an orthopyroxene powder
was used. Dry experimental charge ( 150 mg) along
with 25 ml of water were welded in 5 mm outer dia-
meter platinum capsules. Only fluid-saturated experi-
ments that released water upon recovery are reported in
this study.
ANALYTICAL PROCEDURE
Olivine grains were separated from the host rocks or
removed from the protonation experiments and polished
on both sides to give crystal platelets of 100–600 mm
thickness. The studied crystals were anhedral, so their
crystallographic orientation could not be visually deter-
mined prior to polishing. To have a range of crystal
orientations available, at least 10 crystals per sample
were separated from each rock sample and prepared for
FTIR analysis.
For the FTIR measurements, the crystal platelets were
placed in an IR microscope attached to a Nicolet 800
spectrometer. Spectra were collected in transmission
mode in regions free of cracks or inclusions using an
unpolarized beam. The size of the measuring spot was
defined by choosing either 75 or 100 mm diameter aper-
tures. The IR spectra were collected in the wavenumber
range from 600 to 6000 cm
1
with a resolution of 4 cm
1
.
The IR microscope was kept inside a plastic box and
purged with dried nitrogen gas [see Jamtveit et al. (2001)
for technical details]. The IR spectra were measured after
holding the crystals in a dry nitrogen gas atmosphere for
at least 12 h, resulting in low spectral noise in the wave-
numbers above 3600 cm
1
, where IR stretching of water
vapour occurs (e.g. Bernath, 2002).
ANALYTICAL RESULTS
The FTIR absorption spectra were examined in two
wavenumber ranges: from 3100 to 3700 cm
1
, where
hydroxyl bands absorb; and from 1600 to 2100 cm
1
,
where second-order Si–O overtones occur. The latter
are used to estimate the crystal orientation.
The FTIR spectra exhibit hydroxyl absorption bands
consistent with the Group 1 and Group 2 classification of
Bai & Kohlstedt (1993). Group 1 bands were observed
between 3430 and 3590 cm
1
. The rims of experimen-
tally protonated crystals show a further absorption band
at 3615 cm
1
, which is probably due to high f
H
2
O
in the
experiments. The Group 2 hydroxyl bands occur
between 3285 and 3380 cm
1
. At room temperature,
these bands have FWHM (full width at half maximum)
of 20–40 cm
1
.
Second-order Si–O overtones were used to deduce the
crystallographic orientation of the studied olivine grains
following Jamtveit et al. (2001), but it should be noted that
the directions [100] and [010] were swapped as they
appear to have been wrongly assigned for the commonly
accepted olivine unit-cell settings, where b>c>a.For
several olivine grains, spectroscopically estimated orien-
tations were confirmed by single-crystal X-ray diffracto-
metry (XRD; Fig. 4; Table 4). Below we compare FTIR
spectra that were measured with the IR beam parallel to
[010]. Such spectra are characterized by the strongest
absorption of unpolarized IR radiation in both the
Group 1 and Group 2 wavenumber regions (e.g.
Kohlstedt et al., 1996).
Olivines from Cyprus and Kamchatka
IR absorption spectra of olivine from the Cyprus lavas
are summarized in Fig. 3a. Olivines from orthopyroxene-
bearing samples MAM-25 and TRD-64 absorb in the
Group 2 wavenumber range (spectra 1–4), whereas
absorption in the Group 1 range is at the limit of the
Fig. 2. Correlation between TiO
2
content and SSI of basaltic lavas (&)
and melt inclusions (*). Error bars reflect compositional variations
of melt inclusions in each sample and are calculated at the 95%
confidence level.
608
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
FTIR detection. Olivines from the orthopyroxene-
free Cyprus samples show both the Group 2 and the
Group 1 bands (spectra 5–8). The proportion of
Group 2 bands decreases progressively in the sample
sequence TRD-39, TRD-41–TRD-75–TRD-150, so
that olivine from the tholeiitic sample TRD-150 shows
mainly Group 1 hydroxyl bands (spectrum 8). The FTIR
spectra of the olivine phenocrysts from the Avacha lava
AV-2 contain only Group 1 bands (Fig. 3b).
Placement and intensity of hydroxyl bands in the FTIR
spectra of olivines correlate well with SSI calculated for
basaltic lavas and melt inclusions (Fig. 1). SSI is highest
for the samples in which olivine exhibits Group 2 hydro-
xyl bands and lowest for the samples in which olivine
exhibits mainly Group 1 bands. To better illustrate the
influence of silica saturation on IR absorption, we have
integrated the Group 1 absorption bands from 3466 to
3620 cm
1
and the Group 2 bands from 3260 to
3412 cm
1
, and then determined the absorption intensity
ratio A
int, ratio
¼A
int, Group 1
/(A
int, Group 1
þA
int, Group 2
),
where A
int
denotes integrated absorption coefficients.
The ratio systematically increases with decreasing SSI
of basaltic lavas (Fig. 5a) and melt inclusions (Fig. 5b
and c). Lesser scatter observed for melt inclusions
Table 4: Crystallographic orientation of olivine polished
planes calculated from XRD measurements and data
from Inorganic Crystal Structure Database (ICSD) for
forsteritic olivine (83-1535)
Sample hkl
TRD-64-7 2.26 1.11 0.15
TRD-150-111 5 10 1
TRD-150-9 0 1 0
TRD-150-112 0.31 4.68 0.95
TRD-150-5 0.09 1.33 1.24
TRD-150-6 0.16 0.95 1.34
TRD-64-6 0.29 1.78 2.69
TRD-75-6 0.97 1.36 2.51
We assume unit cell parameters a¼5.990, b¼10.226, c¼
4.769 and Pmnb symmetry group. Calculations were per-
formed using POWD-12þþ software (1997) (Smith et al.,
1982). Samples are listed in the same order as in Fig. 4.
Samples with orientation of polished planes approaching
(100), (010) and (001) are shown in bold in the table, and the
corresponding spectra are highlighted in Fig. 4.
Fig. 3. FTIR absorption spectra of olivine phenocrysts from (a)
Troodos lavas, Cyprus, and (b) a sample from Avacha volcano,
Kamchatka, Russia: 1, 2, MAM-25; 3, 4, TRD-64; 5, TRD-39; 6,
TRD-41; 7, TRD-75; 8, TRD-150; 9, 10, AV-2. Absorption coeffi-
cients are normalized to 1 cm sample thickness and offset to stack the
spectra such that the degree of silica saturation of the parent melt
decreases from top to bottom (see text). The spectra were measured
with an IR beam parallel to [010].
Fig. 4. FTIR spectra of Si–O overtones measured on olivine grains
whose orientation was confirmed by XRD. Bold lines are the spectra
for which the polished planes are close to the principal crystallographic
orientations. The spectra are stacked to illustrate the gradual change in
the overtone spectrum with changing orientation.
609
MATVEEV et al. FTIR SPECTRUM OF OLIVINE
illustrates that melts trapped in olivine preserve the pri-
mary composition better than the host lavas. Based on
the FTIR results obtained for the Cyprus samples, the
transition from the Group 1 dominating spectrum to
the Group 2 dominating spectrum occurs within the
hypersthene-normative field of parental magmas
and in the relatively narrow range of SSI between
06 and 075.
Water solubility in olivines from the Cyprus lavas was
estimated from the Group 1 and Group 2 integrated
absorption coefficients using the calibration of
Libowitzky & Rossman (1997). The weighted mean
wavenumber for the Group 1 hydroxyl bands was located
at 3500 cm
1
and for Group 2 at 3330 cm
1
. Because
FTIR spectra were measured with the unpolarized beam
parallel to [010], an additional orientation factor g¼05
was applied (Mackwell & Kohlstedt, 1990; Lemaire et al.,
2004). Water concentrations are given in Table 5, and for
olivines from the Cyprus lavas fall in the concentration
range from 1 to 2 ppm. However, it should be noted
that the poorly constrained value of orientation factor
(g03–05; Paterson, 1982; Mackwell & Kohlstedt,
1990), spectral noise and unknown olivine matrix correc-
tion of the calibration make these calculations rather
qualitative. Using water concentrations measured on
homogenized melt inclusions in olivine from the Cyprus
lavas as a proxy for water content in the primary melt, an
olivine–melt partition coefficient for H
2
O can be estim-
ated as K
d
(1003) 10
4
. Because the baselines
were positioned at the flanks of each group of absorption
bands, the broad plateau of underlying absorption
commonly attributed to small amounts of molecular
water in submicroscopic inclusions (Miller et al., 1987;
Matveev et al., 2001) did not contribute to the total
calculated water content. Thus water contents reported
here are somewhat lower than those that would result
from integration of the entire hydroxyl absorption area
(e.g. Matveev et al., 2001). The obtained partition coeffi-
cient is in good agreement with the coefficient from Hirth
& Kohlstedt (1996) calculated for shallow mantle and
crustal conditions (<300 MPa), and consistent with a
shallow depth of olivine crystallization.
Hydroxyl in olivine from AV-2 absorbs IR only in
the Group 1 wavenumber range, which is consistent
with the highly silica-undersaturated compositions of
the melt inclusions (Fig. 1b). Low water concentrations
Fig. 5. Integrated absorption coefficient ratio A
int, ratio
¼A
int, Group 1
/
(A
int, Group 1
þA
int, Group 2
) vs SSI of rocks (a) and olivine melt
inclusions (b, c). Vertical error bars are calculated at the 95% con-
fidence level and reflect the compositional range of melt inclusions in
each sample. Error bars in terms of the A
int, ratio
represent the 95%
confidence interval calculated from integrated Group 1 and Group 2
absorbances. These errors include contributions from imperfect IR
beam orientation, spectral noise, and natural variations in A
int, ratio
.
Nepheline-normative and quartz–hypersthene-normative fields are cal-
culated according to the CIPW scheme. Linear regression (dashed line)
calculated for rocks and melt inclusion from Cyprus samples illustrates
the SSI range in which the Group 2 absorption bands replace the
Group 1 hydroxyl absorption bands. The respective compositional
variations are shown in Fig. 1.
610
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
in AV-2 melt inclusions (<01 wt %) might be indi-
cative of higher pressures of olivine crystallization
compared with that of Cyprus olivines, and thus a
somewhat higher olivine–melt H
2
O partition coeffi-
cient (Hirth & Kohlstedt, 1996). Alternatively, water
could have been lost from the melt inclusions during
ascent, as a result of decrepitation (Portnyagin et al.,
2004).
Experimentally protonated olivines from
Solomon Islands picrites
FTIR spectra of experimentally protonated olivine from
sample SS-1 (Mt. Mahimba, Solomon Islands) were
measured from core to rim with 100 mm step incre-
ments using a beam diameter of 100 mm (Fig. 6). Both
experiments produced olivine grains with increased water
contents. The resulting spectra show a 100 mm wide
rim whose defect structure corresponds to the a
SiO
2
of
the buffer used (the lowest spectra in Fig. 6a and b). In
contrast, the cores have lower water contents, but all
show the spectra features related to low silica activity
regardless of the buffer used. Hydroxyl incorporation in
the rim may involve dissolution–precipitation mechan-
isms (e.g. Matveev et al., 2001), whereas core protonation
occurs probably according to reduction–oxidation
reactions as described by Kohlstedt et al. (1997). Good
correlation between FTIR spectra measured on olivine
cores and the low a
SiO
2
implied by the SSI calculated for
SS-1 basalt (Tables 2 and 3, Fig. 5a) implies that the
protonated structure of initially anhydrous olivine can
be used to estimate the degree of silica saturation of the
parent melts.
DISCUSSION
The experiments of Matveev et al. (2001), and the result-
ing FTIR spectra of hydroxyl-bearing mantle olivines,
allow clear discrimination between the effects of a
SiO
2
and other important parameters such as pressure, tem-
perature, water and oxygen fugacities, and olivine major
and trace element compositions. Olivine grains separated
from the same sample at equal P,T,a
H
2
O
, and f
O
2
exhibited Group 1 OH bands when experimentally equi-
librated with magnesiowuustite (low a
SiO
2
) and Group 2
OH bands when equilibrated with orthopyroxene (high
a
SiO
2
). The IR data on olivines from hydrous basaltic
melts obtained in this study support the experimental
results of Matveev et al. (2001), confirming that
natural hydroxyl speciation in olivine is also largely
controlled by a
SiO
2
.
In natural samples, the influence of f
O
2
and trace ele-
ment composition on the FTIR spectrum of olivine may
Table 5: Analytical results
No. of
grains (points)
A
int, ratio
H
2
O (ppm)
MAM-25 4(4) <0.05 1.14 0.18
TRD-64 4(4) <0.05 1.45 0.11
TRD-41 3(4) 0.29 0.06 1.84 0.11
TRD-39 5(4) 0.29 0.04 1.86 0.08
TRD-75 3(2) 0.57 0.05 1.56 0.08
TRD-150 3(3) 0.64 0.07 2.08 0.51
AV-2 2(2) 10
.46 0.1
SS-1 3(6) 1 b.d.l.
Integrated spectra were measured with the IR beam
direction parallel to the [010] crystallographic axis. Errors
are calculated at the 95% confidence level from data
measured on the reported number of grains and points.
b.d.l., below detection limit.
Fig. 6. FTIR spectra of olivine phenocrysts from the Solomon Islands
picrite SS-1, experimentally protonated in periclase (a) and enstatite (b)
matrices. Absorbances are normalized to 1 cm sample thickness. The
lowermost spectra in (a) and (b) are measured on olivine rims, which
equilibrated with the respective a
SiO
2
buffer. All other spectra are from
cores that preserved their pre-experiment defect structures but became
protonated during the experiment. These spectra are taken along
traverses with 100 mm spot intervals.
611
MATVEEV et al. FTIR SPECTRUM OF OLIVINE
often be hard to isolate from the effect of a
SiO
2
and grain
orientation inconsistencies. As f
O
2
controls concentration
of Fe
3þ
in olivine, it may also affect the capacity of olivine
to store hydrogen during natural or experimental proto-
nation (Kohlstedt et al., 1996). In the FTIR spectra of
natural olivines f
O
2
appears to affect the intensities of
individual OH bands (e.g. Matsyuk & Langer, 2004,
and references therein), but because ferric iron is likely
to occur in complexes where it substitutes in neighbour-
ing tetrahedral and octahedral positions (Nakamura &
Schmalzried, 1983) it is unlikely that f
O
2
will significantly
affect the A
int, ratio
.
The effect of trace impurities and the associated extrin-
sic defects in olivine is yet more cryptic. Relative concen-
trations of trace elements were estimated from the melt
inclusion compositions (Table 3) assuming similar
olivine–melt partition coefficients for the variety of the
studied rocks. The A
int, ratio
systematically changes only
with the Ti content of the melt inclusions and thus the
likely Ti content of the olivine (Nikogosian & Sobolev,
1997; Canil & Fedortchouk, 2001). However, the experi-
mental data of Matveev et al., (2001) suggest that the
spectra features assigned to high and low a
SiO
2
can be
reproduced in olivines with identical Ti contents and are,
therefore, not primarily controlled by this parameter.
Thus the apparent correlation between olivine FTIR
spectrum and Ti content is caused by decreasing TiO
2
concentration with increasing a
SiO
2
of the olivine parent
melt (Fig. 2). Nevertheless, the effect of Ti on the olivine
hydroxyl speciation is significant; the data of Berry et al.
(2004) suggest that Ti in olivine strongly affects the exact
position of Group 1 peaks.
Another important effect is f
H
2
O
(P,T,a
H
2
O
), which
affects the solubility of hydroxyl in the olivine structure.
At higher f
H
2
O
not only do the intensities of both Group 1
and Group 2 bands increase, but also the number of
resolvable OH bands, particularly at relatively higher
frequencies (compare the rim and core spectra in Fig. 4;
Kohlstedt et al., 1996; Matveev, 1997; Matveev et al.,
2001; Matsyuk & Langer, 2004). However, there is
no indication that f
H
2
O
notably affects the A
int, ratio
(Matveev, 1997; Matveev et al., 2001).
Considering a
SiO
2
as a key variable controlling the fre-
quency of OH IR absorption in olivine, we suggest that
the FTIR spectrum of olivine can be used to deduce the
a
SiO
2
at crystallization or final equilibration. The good cor-
relation between IR spectra and the composition of melt
inclusions or host lavas implies that the defect structure of
the studied olivines has survived changes in pressure–
temperature and even a
SiO
2
conditions during ascent.
CONCLUSIONS
The FTIR spectra of olivine phenocrysts in basaltic lavas
correlate well with the degree of silica saturation of the
parent melt and therefore a
SiO
2
. Liquidus olivines crystal-
lized from nepheline-normative basaltic melts have OH
bands between 3430 and 3590 cm
1
(Group 1). Olivines
that coexist at magmatic temperatures with orthopyrox-
ene have OH bands in the wavenumber range from 3285
to 3380 cm
1
(Group 2). Olivines from orthopyroxene-
undersaturated hypersthene-normative basaltic melts
exhibit both groups of hydroxyl absorption bands, with
the proportion of Group 2 bands increasing with increas-
ing SSI. The compositions of melt inclusions correlate
better with the FTIR spectra of olivine than the com-
position of the host basaltic lava, and thus more
accurately preserve information on the silica saturation
of primary melts.
FTIR results from this study are consistent with the
experimental results of Matveev et al. (2001), implying
that natural hydroxyl occurrence in olivine is similar to
that imposed on natural olivine in high-pressure experi-
ments. Therefore spectroscopically ‘dry’ olivine defect
structures that have lost hydrogen during magma ascent
can be re-protonated experimentally to reveal the a
SiO
2
at
their last vacancy equilibration. Hydrogen diffusion and
reduction of ferric iron in the olivine structure to OH and
ferrous iron are relatively rapid, such that the original
defect structure may survive a short experimental run
time, and may be made visible by subsequent IR spectro-
scopy to reveal the original a
SiO
2
conditions.
ACKNOWLEDGEMENTS
We thank the lapidary workshop at Muunster University
for sample preparation. Financial support by the DFG
(Deutsche Forschungsgemeinschaft) through grants Ba
964/16-1 and Ge 659/11-1 (to C.B. and C.G.), the
European Commission IHP Programme grant, which
allowed IR analyses at the University of Bristol (to
S.M.), as well as BMBF (Bundesministerium fuur Bildung
und Forschung) funded KOMEX-2 project and RFBR
(Russian Foundation for Basic Research) through grant
03-0564629 (to M.P.) are gratefully acknowledged. Ear-
lier discussions with A. Sobolev (MPI fuur Chemie, Mainz)
were invaluable for structuring the study. We also thank
R. W. Luth, T. Chacko (University of Alberta) and
J. Harris (University of Glasgow) for their comments,
which helped to improve the manuscript. We thank
J. Loens (Muunster University) for performing single-
grain XRD analyses.
REFERENCES
Aines, R. D. & Rossman, G. R. (1984). Water in minerals? A peak in
the infrared. Journal of Geophysical Research 89, 4059–4071.
Bai, Q. & Kohlstedt, D. L. (1993). Effects of chemical environment on
the solubility and incorporation mechanism for hydrogen in olivine.
Physics and Chemistry of Minerals 19, 460–471.
612
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
Beran, A. (1970). U
¨ber (OH)-Gruppen in Olivin. (The OH-groups in
olivine.) Anzeiger der
OO
sterreichischen Akademie der Wissenschaften,
Mathematisch Naturwissenschaftliche Klasse 106, 73–74.
Beran, A. & Putnis, A. (1983). A model of the OH positions in olivine,
derived from infrared-spectroscopic investigations. Physics and
Chemistry of Minerals 9, 57–60.
Bernath, P. F. (2002). The spectroscopy of water vapour: experiment,
theory and applications. Physical Chemistry and Chemical Physics 4,
1501–1509.
Berry, A., Hermann, H. & O’Neill, H. (2004). The water site in mantle
olivine. Geochimica et Cosmochimica Acta 68, A36.
Braithwaite, J. S., Wright, K. & Catlow, C. R. A. (2003). A theoretical
study of the energetics and IR frequencies of hydroxyl defects in
forsterite. Journal of Geophysical Research 108, ECV 1–ECV 9.
Cameron, W. E. (1985). Petrology and origin of primitive lavas from
the Troodos ophiolite, Cyprus. Contributions to Mineralogy and Petrology
89, 239–255.
Canil, D. & Fedortchouk, Y. (2001). Olivine–liquid partitioning of
vanadium and other trace elements; with applications to modern
and ancient picrites. Canadian Mineralogist 39, 319–330.
Carmichael, I. S. E. (2002). The andesite aqueduct; perspectives on the
evolution of intermediate magmatism in west–central (105–99
degrees W ) Mexico. Contributions to Mineralogy and Petrology 143,
641–663.
Danyushevsky, L., McNeill, A. W. & Sobolev, A. V. (2002).
Experimental and petrological studies of melt inclusions in
phenocrysts from mantle-derived magmas: an overview of techni-
ques, advantages and complications. Chemical Geology 183, 5–24.
Demouchy, S. & Mackwell, S. (2003). Water diffusion in synthetic iron-
free forsterite. Physics and Chemistry of Minerals 30, 486–494.
Falloon, T. J. & Danyushevsky, L. V. (2000). Melting of refractory
mantle at 15, 2 and 25 GPa under anhydrous and H
2
O-
undersaturated conditions; implications for the petrogenesis of
high-Ca boninites and the influence of subduction components on
mantle melting. Journal of Petrology 41, 257–283.
Falloon, T. J. & Green, D. H. (1987). Anhydrous partial melting of
MORB pyrolite and other peridotite compositions at 10 kbar:
implication for the origin of primitive MORB glasses. Mineralogy and
Petrology 37, 181–219.
Hirth, G. & Kohlstedt, D. L. (1996). Water in the oceanic upper
mantle: implications for rheology, melt extraction and the evolution
of the lithosphere. Earth and Planetary Science Letters 144, 93–108.
Ingrin, J. & Skogby, H. (2000). Hydrogen in nominally anhydrous
upper-mantle minerals; concentrations levels and implications.
European Journal of Minerals 12, 543–570.
Jamtveit, B., Brooker, R., Brooks, K., Larsen, L. M. & Pedersen, T.
(2001). The water content of olivines from the North
Atlantic Volcanic Province. Earth and Planetary Science Letters 186,
401–415.
Kohlstedt, D. L. & Mackwell, S. J. (1998). Diffusion of hydrogen and
intrinsic point defects in olivine. Zeitschrift fuur Physikalische Chemie—
International Journal of Research in Physical Chemistry and Chemical Physics
2, 147–162.
Kohlstedt, D. L., Keppler, H. & Rubie, D. C. (1996). Solubility of
water in the alpha, beta and gamma phases of (Mg,Fe)
2
SiO
4
.
Contributions to Mineralogy and Petrology 123, 345–357.
Kohn, S. C. (1996). Solubility of H
2
O in nominally anhydrous
mantle minerals using H-1 MAS NMR. American Mineralogist 81,
1523–1526.
Lemaire, C., Kohn, S. C. & Brooker, R. A. (2004). The effect of silica
activity on the incorporation mechanisms of water in synthetic
forsterite: a polarised infrared spectroscopic study. Contributions to
Mineralogy and Petrology 147, 48–57.
Libowitzky, E. & Rossman, G. R. (1997). An IR absorption calibration
for water in minerals. American Mineralogist 82, 1111–1115.
Mackwell, S. J. & Kohlstedt, D. L. (1990). Diffusion of hydrogen in
olivine: implications for water in the mantle. Journal of Geophysical
Research 95, 5079–5088.
Mackwell, S. J., Dimos, D. & Kohlstedt, D. L. (1988). Transient creep
of olivine: point-defect relaxation times. Philosophical Magazine A 57,
779–789.
Matsyuk, S. S. & Langer, K. (2004). Hydroxyl in olivines from mantle
xenoliths in kimberlites of the Siberian platform. Contributions to
Mineralogy and Petrology 147, 413–437.
Matveev, S. (1997). CHO fluids at upper mantle conditions. Ph.D.
thesis, Universitaat zu Kooln, Cologne.
Matveev, S., O’Neill, H. S. C., Ballhaus, C., Taylor, W. R. &
Green, D. H. (2001). Effect of silica activity on OH
IR spectra of
olivine: implications for low-a
SiO
2
mantle metasomatism. Journal of
Petrology 42, 721–729.
Miller, G. H., Rossman, G. R. & Harlow, G. E. (1987). The natural
occurrence of hydroxide in olivine. Physics and Chemistry of Minerals 14,
461–472.
Nakamura, A. & Schmalzried, H. (1983). On the nonstoichiometry
and point defects of olivine. Physics and Chemistry of Minerals 10,
27–37.
Nikogosian, I. K. & Sobolev, A. V. (1997). Ion-microprobe analysis of
melt inclusions in olivine: experience in estimating the olivine–melt
partition coefficients of trace elements. Geochemistry International 35,
119–126.
Paterson, M. (1982). The determination of hydroxyl by infrared
absorption in quartz, silicate glasses and similar materials. Bulletin de
Mineeralogie 105, 20–29.
Pearce, J. A. (1975). Basalt geochemistry used to investigate past
tectonic environments on Cyprus. Tectonophysics 25, 41–67.
Portnyagin, M. V. (1998). Origin of mantle derived magmas in supra-
subduction zones: a case study of the Troodos Ophiolite (Cyprus).
Ph.D. thesis, Vernadski Institute, Moscow, 230 pp. (in Russian).
Portnyagin, M. V., Magakyan, R. & Schminke, H. U. (1996).
Geochemical variability of boninite magmas; evidence from
magmatic inclusions in high magnesian olivine from lavas of
southwestern Cyprus. Petrologiya 4, 250–265.
Portnyagin, M. V., Danyushevsky, L. V. & Kamenetsky, V. S. (1997).
Coexistence of two distinct mantle sources during formation of
ophiolites: a case study of primitive pillow-lavas from the lowest part
of the volcanic section of the Troodos Ophiolite, Cyprus.
Contributions to Mineralogy and Petrology 128, 287–301.
Portnyagin, M. V., Plechov, P. Yu., Matveev, S. V. & Mironov, N. L.
(2004). Petrology of the ‘avachites’— high-magnesian basalts from
Avachinsky volcano (Kamchatka): II. Melt inclusions in olivine.
Petrology (in press).
Rautenschlein, M., Jenner, G. A., Hertogen, J., Hofmann, A. W.,
Kerrich, R., Schmincke, H. U. & White, W. M. (1985). Isotopic and
trace element composition of volcanic glasses from the Akaki
Canyon, Cyprus; implications for the origin of the Troodos
Ophiolite. Earth and Planetary Science Letters 75, 369–383.
Rohrbach, A. (2003). Petrologische Untersuchungen an Pikriten der
Salomonen-Inseln. Diploma thesis, Westfaalische Wilhelms
Universitaat, Muunster, 61 pp.
Shervais, J. W. (2001). Birth, death, and resurrection: the life cycle of
suprasubduction zone ophiolites. Geochemistry, Geophysics, Geosystems,
45, paper number 2000GC000080.
Smith, D. K., Nichols, M. C. & Zolensky, M. E. (1982). POWD10, a
FORTRAN IV program for calculating X-ray powder diffraction
patterns—Version 10. University Park: Pennsylvania State
University, 103 pp.
613
MATVEEV et al. FTIR SPECTRUM OF OLIVINE
Sobolev, A. V. & Chaussidon, M. (1996). H
2
O concentrations in
primary melts from supra-subduction zones and mid-ocean ridges:
implications for H
2
O storage and recycling in the mantle. Earth and
Planetary Science Letters 137, 45–55.
Sobolev, A. V., Portnyagin, M. V., Dmitriyev, L. V., Tsameryan, O. P.,
Danyushevskiy, L. V., Kononkova, N. N., Shimizu, N. & Robinson, P.
(1993). Petrology of ultramafic magmas and associated rocks of the
Troodos Massif, Cyprus. Petrologiya 1, 379–412 (in Russian).
Stocker, R. L. & Smyth, D. M. (1978). Effect of enstatite activity and
oxygen partial pressure on the point-defect chemistry of olivine.
Physics of the Earth and Planetary Interiors 16, 146–156.
Wilkins, R. W. T. & Sabine, W. (1973). Water content of some
nominally anhydrous silicates. American Mineralogist 58, 508–516.
Wright, K. & Catlow, C. R. A. (1994). A computer simulation study
of (OH) defects in olivine. Physics and Chemistry of Minerals 20,
515–518.
614
JOURNAL OF PETROLOGY VOLUME 46 NUMBER 3 MARCH 2005
... The interest in hydrogen as an important chemical component of Earth's mantle escalated when hydroxyl groups were observed in nominally anhydrous minerals (NAMs) such as (Mg,Fe) 2 SiO 4 olivine using Fourier Transform Infrared Spectroscopy, [1][2][3][4][5][6][7][8][9], leading to the idea that H 2 O may not only be present but have a profound influence on the dynamics of the Earth even at great depths. More recently, Pearson et al. [10] observed significant H 2 O concentrations in natural crystals of HP and HT polymorphs of olivine preserved within diamond inclusions, giving further confirmation that water is present at great depths in the innermost regions of the Earth. ...
... The first, and most common, mechanism involves Si vacancies (in Kröger-Vink notation, (Si) = (4H) x Si ), where four protons completely fill a tetrahedral vacancy, each bonded to one of the apical oxygens of the tetrahedron. This mechanism leads to a final stoichiometry of Mg 2 H 4 O 4 and is commonly called a "hydrogarnet defect" [4,5,[48][49][50]. The major FTIR absorption peaks associated with this defect usually occur in the higher frequency region where OH stretching vibrations are observed, between 3400 and 3700 cm −1 . ...
... Peaks identified at 3476, 3536, 3550, 3566, 3578, 3605, and 3612 cm −1 are related to a specific mechanism of hydrogen incorporation in forsterite called the "hydrogarnet substitution" [4,5,48,50]. This mechanism involves Si vacancies, where four protons completely fill a tetrahedral vacancy, each bonded to one of the apical oxygens of the tetrahedron, leading to a final stoichiometry of Mg 2 H 4 O 4 . ...
Article
Full-text available
Water distribution in the deep Earth represents one of the most important topics in the field of geodynamics due to its large impact on the physical and chemical properties of the Earth's mantle, such as electrical conductivity, seismic anisotropy, diffusion, and rheology. In this study, we synthesized hydrous forsterite at 1100 • C and up to 4 GPa with either a piston-cylinder or multianvil apparatus. As a starting material, we used synthetic forsterite, unbuffered by SiO 2 , obtained by thermo-mechanical activation of talc and magnesium carbonate hydroxide. Hydration was carried out using liquid H 2 O as hydrogen source. Samples were polycrystalline in an effort to distribute H 2 O throughout the sample both rapidly and homogeneously. Using the Paterson calibration, we observed total water content concentrations ranging between 100 and 500 ppm wt H 2 O. Multiple absorption bands are found in the frequency range between 3400 and 3650 cm −1 , identifying at least seven peaks in all samples. Vibrational bands were centered at 3476, 3535, 3550, 3566, 3578, 3605, and 3612 cm −1 , in good agreement with experimental studies conducted on both hydrous forsterite and single crystals of olivine. The stronger OH stretching peaks can be attributed to vibrational modes associated with the hydrogarnet defect 4H x Si in which four protons occupy a vacant tetrahedral site. None of the OH bands observed are found at frequencies associated with hydrogen occupying vacant octahedral sites. High-temperature FTIR spectroscopy was used to evaluate the evolution of IR spectra as a function of temperature, up to 500 • C. The complete reversibility of peak absorption vs. temperature in the OH stretching region confirms that no water loss occurred during heating. We observe an overall a decrease in total absorption with increasing temperature, and a prominent decrease in the relative intensities of the higher frequency bands (>3600 cm −1) with respect to lower frequency bands. We have assigned a series of equilibrium expressions based on the variation of relative peak areas with temperature and find that enthalpies of these processes range between 0.047-0.068 eV (4.5-6.5 kJ/mol), very low in comparison to activation energies observed for electrical conduction in hydrous olivine. Major changes in the vibrational spectrum are expected to be related to configurational changes of the same fully protonated hydrogarnet defect species. However, the complexity of the FTIR spectra may also be related to partially protonated defects, such as the associate defect 3H Si + H • i generated by a dissociation reaction of the hydrogarnet species.
... The interest in hydrogen as an important chemical component of Earth's mantle escalated when hydroxyl groups were observed in nominally anhydrous minerals (NAMs) such as (Mg,Fe)2SiO4 olivine using Fourier Transform Infrared Spectroscopy, [1][2][3][4][5][6][7][8][9], leading to the idea that H2O may not only be present but have a profound influence on the dynamics of the Earth even at great depths. More recently, Pearson et al. [10] observed significant H2O concentrations in natural crystals of HP and HT polymorphs of olivine preserved within diamond inclusions, giving further confirmation that water is present at great depths in the innermost regions of the Earth. ...
... The first, and most common, mechanism involves Si vacancies (in Kröger-Vink notation, (Si) = (4H) Si x ), where four protons completely fill a tetrahedral vacancy, each bonded to one of the apical oxygens of the tetrahedron. This mechanism leads to a final stoichiometry of Mg2H4O4 and is commonly called a "hydrogarnet defect" [4,5,[48][49][50]. The major FTIR absorption peaks associated with this defect usually occur in the higher frequency region where OH stretching vibrations are observed, between 3400 and 3700 cm −1 . ...
... Peaks identified at 3476, 3536, 3550, 3566, 3578, 3605, and 3612 cm −1 are related to a specific mechanism of hydrogen incorporation in forsterite called the "hydrogarnet substitution" [4,5,48,50]. This mechanism involves Si vacancies, where four protons completely fill a tetrahedral vacancy, each bonded to one of the apical oxygens of the tetrahedron, leading to a final stoichiometry of Mg2H4O4. ...
Article
Full-text available
Water distribution in the deep Earth represents one of the most important topics in the field of geodynamics due to its large impact on the physical and chemical properties of the Earth’s mantle, such as electrical conductivity, seismic anisotropy, diffusion, and rheology. In this study, we synthesized hydrous forsterite at 1100 C° and up to 4 GPa with either a piston-cylinder or multianvil apparatus. As a starting material, we used synthetic forsterite, unbuffered by SiO2, obtained by thermo‐mechanical activation of talc and magnesium carbonate hydroxide. Hydration was carried out using liquid H2O as hydrogen source. Samples were polycrystalline in an effort to distribute H2O throughout the sample both rapidly and homogeneously. Using the Paterson calibration, we observed total water content concentrations ranging between 100 and 500 ppm wt H2O. Multiple absorption bands are found in the frequency range between 3400 and 3650 cm−1, identifying at least seven peaks in all samples. Vibrational bands were centered at 3476, 3535, 3550, 3566, 3578, 3605, and 3612 cm−1, in good agreement with experimental studies conducted on both hydrous forsterite and single crystals of olivine. The stronger OH stretching peaks can be attributed to vibrational modes associated with the hydrogarnet defect 4HSix in which four protons occupy a vacant tetrahedral site. None of the OH bands observed are found at frequencies associated with hydrogen occupying vacant octahedral sites. High-temperature FTIR spectroscopy was used to evaluate the evolution of IR spectra as a function of temperature, up to 500 °C. The complete reversibility of peak absorption vs temperature in the OH stretching region confirms that no water loss occurred during heating. We observe an overall a decrease in total absorption with increasing temperature, and a prominent decrease in the relative intensities of the higher frequency bands (>3600 cm−1) with respect to lower frequency bands. We have assigned a series of equilibrium expressions based on the variation of relative peak areas with temperature and find that enthalpies of these processes range between 0.047–0.068 eV (4.5–6.5 kJ/mol), very low in comparison to activation energies observed for electrical conduction in hydrous olivine. Major changes in the vibrational spectrum are expected to be related to configurational changes of the same fully protonated hydrogarnet defect species. However, the complexity of the FTIR spectra may also be related to partially protonated defects, such as the associate defect 3HSi’ + Hi• generated by a dissociation reaction of the hydrogarnet species.
... The ability to dissolve hydrogen as OH groups makes olivine the main water reservoir in the upper mantle and an important source of information about the mantle fluid regime. The water content in olivine is controlled by the temperature, pressure [1], and the activity of H 2 O, MgO, SiO 2 , and FeO in its crystallization environment [2][3][4][5][6]. At present, there are no clear concepts concerning the solubility of water in olivine formed at pressures, temperatures, and low oxygen fugacity ( fO 2 ) typical of the base of the subcratonic lithosphere [7]. ...
... In equilibrium with the hydrocarbon fluid, olivine contains much less water when in equilibrium with an aqueous fluid at a similar P-T-fH 2 . In hydrous mantle olivines, OH peaks were found only in the high-frequency region [2]. This allowed us to conclude that the water-rich samples of mantle olivine were not in equilibrium with the reduced fluid consisting of light alkanes or oxygenated hydrocarbons. ...
Article
The main trends of water dissolution in Fe-bearing olivine have been investigated in the olivine-H 2 O-hydrocarbon fluid system in experiments at a pressure of 6.3 GPa, a temperature of 1200°C, and hydrogen fugacity (fH 2) buffered by the Mo-MoO 2 equilibrium. The content and contribution of ОH defects of different types in Fe-bearing olivines depend on the composition of reduced fluids in the system. As the fraction of hydrocarbons in the fluid increases, the H 2 O content in olivine crystals decreases from 900 to 160-180 ppm, while the ОН absorption peaks become lower at high frequencies and occupy a larger part of the infrared spectrum in the low-frequency region. According to the experimental results, even the deepest seated mantle olivines with OH defects were not equilibrated with a fluid rich in light alkanes or oxygenated hydrocarbons .
... In particular, for the Earth's mantle, it is olivine whose H 2 O storage capacity defines the position of the lithosphere-asthenosphere boundary [5]. The water content in olivine is governed by the pressure and temperature conditions and the activities of H 2 O, MgO, SiO 2 , and FeO in the crystallization environment [6][7][8][9][10][11][12][13][14]. A specific set of these parameters, characteristic of the mantle rocks of the terrestrial planets, determines the H 2 O storage capacity of olivine. ...
Article
Full-text available
The key features of the interaction between peridotites of the continental lithospheric mantle and reduced hydrocarbon-rich fluids have been studied in experiments conducted at 5.5 GPa and 1200 °C. Under this interaction, the original harzburgite undergoes recrystallization while the composition of the fluid changes from CH4-H2O to H2O-rich with a small amount of CO2. The oxygen fugacity in the experiments varied from the iron-wustite (IW) to enstatite-magnesite-olivine-graphite/diamond (EMOG) buffers. Olivines recrystallized in the interaction between harzburgite and a fluid generated by the decomposition of stearic acid contain inclusions composed of graphite and methane with traces of ethane and hydrogen. The water content of such olivines slightly exceeds that of the original harzburgite. Redox metasomatism, which involves the oxidation of hydrocarbons in the fluid by reaction with magnesite-bearing peridotite, leads to the appearance of additional OH absorption bands in the infrared spectra of olivines. The water content of olivine in this case increases by approximately two times, reaching 160–180 wt. ppm. When hydrocarbons are oxidized by interaction with hematite-bearing peridotite, olivine captures Ca-Mg-Fe carbonates, which are products of carbonate melt quenching. This oxidative metasomatism is characterized by the appearance of specific OH absorption bands and a significant increase in the total water content in olivine of up to 500–600 wt. ppm. These findings contribute to the development of criteria for reconstructing metasomatic transformations in mantle rocks based on the infrared spectra and water content of olivines.
... Meanwhile, the stretching vibration at 2937 cm -1 and 2895 cm -1 belonging to C-H 2 [32,33]. The vibration at 1660 cm -1 , 1200 cm -1 and 1086 cm -1 are allocated to C=C stretching, and C-O stretching [34,35], respectively with higher intensity in the PVA-G-Ag nanocomposite spectrum than in pure PVA and graphite spectrums, indicating that carbon bond between C, C and O has been established. The above results prove the strong interfacial interaction between graphene and PVA. ...
Conference Paper
Full-text available
The PVA-G-Ag nanocomposite have been synthesized effectively by pulsed laser ablation liquid (PLAL) as a considered to be environmentally friendly and free of residues from chemical reactions. The high excellence silver plate (99.99%) and graphite plate (99.99%) was immersed in the polyvinyl alcohol (PVA) solution and irradiated with the Nd-YAG laser at wavelength 1064 nm, power 160 mJ for the silver plate and 80mJ for graphite plate, reiteration rate 6 Hz, 10 ns pulse width and 300 pules for graphite plate and 700 pulse for silver plate. The pure of PVA, PVA-Graphene and PVA-Graphene-Ag nanocomposite were investigated using UV-VIS spectroscopy, FTIR and SEM. The absorption spectra of PVA-Graphene-Ag nanocomposite show the presence of two peaks one 0.4 at 272 and second 0.47 at 403 nm. The optical energy gap (Eg) decreased from 5eV of a pure PVA to 4.6eV of a PVA-G-Ag for indirect allowed transition and therefore, decreased from 4.4eV of a pure PVA to 4.1eV of a PVA-G-Ag for indirect forbidden transition. The transmittance and absorption coefficient have been determined. The SEM images confirmed that homogenous composite without aggregation of the components. The average size of nanoparticles of GNPs and AgNPs for PVA-G and PVA-G-Ag nanocomposite was 130 and 115 nm respectively. The FTIR has demonstrated that the connection between the graphene, silver and polymer network was enough to have stable nanocomposite. This investigation demonstrates that the pulse laser ablation decent instrument to decorated metals on the graphene with the presence of the polymer.
... The group I OH bands (Bai and Kohlstedt 1992) are dominant in this study. Although some earlier studies (e.g., Matveev et al. 2001Matveev et al. , 2004Lemaire et al. 2004) suggested that the group I bands should not dominate at high silica activity buffered sample, later studies (e.g., Smyth et al. 2006) suggested that silica activity has minimal effect on the sites of H incorporation at high pressures. In addition, the band positions in this study are consistent with those in previous studies that performed at similar thermodynamic conditions (e.g., Kohlstedt et al. 1996;Mosenfelder et al. 2006;Kovacs et al. 2010) within the resolution of the spectrometer. ...
Article
Full-text available
A transition from A-type to E-type fabrics in olivine may be the cause of a decrease in seismic anisotropy with depth in the upper mantle. To better understand upper mantle seismic signals, we investigate the origin of E-type fabrics using a natural olivine by deformation experiments. An olivine crystal was first hydrated at 5 GPa and 1473 K (with 4-60 ppm H2O), or dehydrated at room pressure at 1473 K at an oxygen fugacity near the enstatite-magnesite-olivine-graphite (EMOG) buffer. This hydrated/dehydrated olivine was then sheared in the [100] direction on the (001) plane at pressures of 2 to 5 GPa and temperatures of 1473 or 1573 K. The deformed samples were observed by transmission electron microscopy (TEM) on the (001) plane to determine whether the [100](001) slip system was activated or not. Only c-elongated [100] dislocations were observed for the anhydrous samples, while [100](001) dislocations dominated in the hydrous samples. The dislocation structure of the [100](001) slip system developed under hydrous and relatively low-temperature conditions indicates different slip mechanism which is detected under anhydrous and high-temperature conditions in previous studies. We conclude that the incorporation of water into olivine helps to activate the [100](001) slip system by reducing its Peierls stress. This supports the idea that E-type fabrics can exist under hydrous conditions and that a transition to this fabric may be the cause of seismic anisotropy decrease with depth in the asthenosphere.
Article
Due to the multilayered structure of the skin tissue, the architecture of its engineered scaffolds needs to be improved. In the present study, 45s5 bioglass nanoparticles were selected to induce fibroblast proliferation and their protein secretion involved in wound healing, although cobalt ions were dopped to increase their potency. 3-layer scaffold was designed and fabricated via electrospinning method: polyurethane (PU) as outer layer, polycaprolactone (PCL) /collagen as middle layer and BGs nanoparticle/PCL/collagen as inner layer. Two three-layer scaffold with difference in inner layer fabricate and classified in two experimental groups. The scaffold groups were then examined in the following by scanning electron microscopy (SEM) and fourier transform infrared (FTIR) methods. Also, the mechanical assay showed the highest young modulus for the composite scaffold with the doped nanoparticles and the water contact angle of this scaffold after chemical crosslinking of collagen was reduced to 52.34 ± 7.7°. On the other hand, the weight loss of the corresponding scaffold was the highest value of 82.35 ± 4.3 % due to the alkaline effect of metal ions. Moreover, better cell expansion, greater cell confluence and a lower degree of toxicity were confirmed by using L-929 fibroblasts for the composite scaffold with the substituted nanoparticles. Also, the up-regulation of TGF β1 and VEGF introduced this composite scaffold as a better model for the fibroblasts commitment to a new skin tissue without scar. As a conclusion, the 3-layered scaffold which is loaded with cobalt ions-bonded bioglass nanoparticles, is a better substrate for the culture of the fibroblasts and dermal regeneration.
Article
It is believed that the Moon formed following collision of a large planetesimal with the early Earth. Over the ∼4 Gyr since this event the Moon has been considerably less processed by geological activity than the Earth, and may provide a better record of processes and conditions in the early Earth-Moon system. There have been many studies of magmatic volatiles such as H, F, Cl, S and C in lunar materials. However, our ability to interpret variable volatile contents in the lunar sample suite is dependent on our understanding of volatile behaviour in lunar systems. This is currently constrained by limited experimental data. Here, we present the first experimental mineral-melt partitioning coefficients for F, Cl and H2O in a model lunar system under appropriately reduced conditions (log fO2 to IW-2.1, i.e. oxygen fugacity down to 2.1 log units below the Fe-FeO buffer). Data are consistent with structural incorporation of F, Cl and OH⁻ in silicate melt, olivine and pyroxene under conditions of the lunar mantle. Oxygen fugacity has a limited effect on H2O speciation, and partitioning of H2O, F and Cl is instead largely dependent on mineral chemistry and melt structure. Partition coefficients are broadly consistent with a mantle source region for lunar volcanic products that is significantly depleted in F, Cl and H2O, and depleted in Cl relative to F and H2O, compared to the terrestrial mantle. Partitioning data are also used to model volatile redistribution during lunar magma ocean (LMO) crystallisation. The volatile content of lunar mantle cumulates is dependent upon proportion of trapped liquid during LMO solidification. However, differences in mineral-melt partitioning during LMO solidification can result in significant enrichment on F relative to Cl, and F relative to H2O, in cumulate phases relative to original LMO composition. As such, Cl depletion in lunar volcanic products may in part be a result of LMO solidification.
Article
Recent studies of minor elements (CaO, Al2O3 , MnO, NiO, P2O5 and H2O) in igneous olivines are reviewed. CaO content is used to distinguish between igneous and peridotitic olivines because CaO contents of igneous and peridotitic olivines are typically higher and lower than ~0.1 wt.%, respectively. However, recent studies suggest that olivine-melt Ca partition coefficient decreases as melt H2O content increases, and as a result, arc magmas sometimes crystallize olivines with CaO content comparable to those of peridotitic olivines. Therefore, it is not straightforward to distinguish between igneous and peridotitic olivines in arc magmas based on only CaO content. Aluminum content of olivine in equilibrium with spinel increases as temperature increases; the relation is formulated as Al-in-olivine geothermometer, which enables us to estimate crystallization temperatures of igneous olivines. Nickel contents and Mn/Fe ratios of olivines are used to identify the source lithology of their parental magmas; Ni content and Mn/Fe ratio are thought to increase and decrease with increasing a proportion of pyroxenite-derived component in the parental melt, respectively. However, recent experimental and petrological studies contest the validity of this notion. Phosphorus concentration mapping is used to clarify crystal growth kinetics and diffusion processes of igneous olivines. The results suggest that dendritic growth is a common growth mechanism of igneous olivines. Hydrogen contents in igneous olivines are 10-100 times lower than those of peridotite-derived olivine xenocrysts. Hydrogen diffusion is significantly rapid in olivine, and therefore H2O contents of olivine-hosted melt inclusions reequilibrate with the olivine-hosting magmas in short time. Minor elements in igneous olivines offer constraints on pre-eruptive magmatic conditions and crystallization process, but extracting information about physicochemical conditions and source lithology of their parental melts is not straightforward.
Article
Full-text available
Detailed geochemical, mineralogical, and experimental studies of an Ol-phyric boninite from the Mavrokolimbos area, southwestern Cyprus are presented. Olivine phenocrysts have a primitive composition (Fo87-92). They contain crystalline inclusions of magnesian orthopyroxene (Mg# = 0.90-0.92), chromian spinel of variable composition [Cr/(Cr + Al) = 0.79-0.60], and syngenetic melt and fluid inclusions. The melt inclusions, which were quenched experimentally, were analyzed for major elements (using electron microprobe), trace elements, and H2O (by ion microprobe). The detected compositional variations of the melt inclusions suggest a wide chemical diversity of the primary boninitic magmas and the involvement of at least four independent components in their composition. According to the proposed model, the boninite melts formed by partial mantle melting in an open system, at an effective restite permeability of approximately 2%, a fact evident from the occurrence of an ultradepleted melt [(CeTYb)n = 0.07]. The incipient mantle melting began under the effect of the aqueous component. The primary fluid was characterized by a low content of the most incompatible elements (H2O/K2O > 150). Its origin was related to the dehydration of a subducted slab. The enrichment of the boninite in LREE and LILE was due to the presence of an independent component [(La/Sm)n > 2, (K/La)n ≈ 5, (K/Sr)n ≈ 5, (Sr/Ce)n ≈ 1], which could be introduced in the mantle source by the aqueous component percolating through the heterogeneously metasomatized mantle. Alternatively, the enriched component may have locally occurred in the mantle in the form of melt films before the melting. The genesis of this component is related to potassic mantle metasomatism. The boninitic magmas interacted with silicic rocks/melts during the crystallization and produced hybrid melts with anomalously high contents of HREE [Yb > 2.5 ppm, (Dy/Ti)n > 5], Y, and Na, and low CaO/Al2O3 ratios (<0.75). Copyright © 1996 by MAEe Cyrillic signK Hayκa/Interperiodica Publishing.
Article
Full-text available
The composition and crystallization conditions of the parental melts of avachites were elucidated by studying melt inclusions in olivine (Fo 85.8-90.7) phenocrysts. The melt inclusions captured during the crystallization of primitive magmas subsequently reequilibrated with their host minerals and became partly recrystallized and decrepitated. The diffusion-controlled reequilibration of the melt inclusions with the olivine occurred at temperatures close to ∼1100°C and was associated with the crystallization of daughter phases: olivine, high-Ca pyroxene, and spinel. The composition of the pyroxene and spinel in the inclusions evolved toward extremely high Al contents, which is atypical of pyroxene in the rocks and was controlled by plagioclase absence from the daughter phase assemblage of the inclusions. Magma decompression induced the partial decrepitation of the melt inclusions, a process that was associated with the escape of fluid components (CO2 and H2O) and variable amounts of the residual silicate material from the inclusions. The initial compositions of the melt inclusions, which were reconstructed using techniques of experimental homogenization and modeling, show broad ranges in the contents of major and trace elements. Compared with the composition of the rocks, the compositions of inclusions in the olivine Fo > 90% are higher in CaO, A12O 3, and Na2O at lower concentrations of SiO2. Their geochemical characteristics are identical to those of low-Si ankaramite melts occurring in many island arcs. The carbonatite metasomatism of the arc mantle, the derivation of nepheline-normative ankaramite magmas, and the significant crustal contamination of these magmas during their fractionation can be spread more widely than is currently assumed in models for island-arc petrogenesis. The evolution of the avachite primitive magmas was controlled by the crystallization of early olivine, high-Ca pyroxene, spinel, and, perhaps, the assimilation of crustal rocks in the magmatic chambers at different depths (from 5 to 30 km). During two (or more) crystallization stages, olivine-pyroxene cumulates were produced, remobilized, and transported to the surface by the differentiated hypersthene-normative magmas. Avachites are hybrid cumulative rocks, which were produced in a long-lived open magmatic system.
Article
Full-text available
Primary melt inclusions in olivine from the picrite of Mauna Loa volcano, Hawaii Island, were studied by the methods of high-temperature optical thermometry, secondary ion mass spectrometry, and X-ray electron microprobe analysis to quantify the partitioning coefficients (Kd) of Ti, Dy, Li, Y, Yb, and Cr between magnesian olivine and tholeiitic melt at temperatures of 1330-1280°C, a pressure of @ca. 1 kbar,fO2 at approximately the QFM buffer, and element contents approximating those in natural rocks. A significant linear correlation between In Kd and 1/T was established for Ti, Li, and Cr. The correspondence between the Kd values determined for Ti, Dy, Y, Yb, and Cr with the published ranges of these values indicates that studies of melt inclusions in minerals offer great promise as a method for determining Kd of elements between minerals and melts.
Article
We have analyzed the kinetics of hydrogen diffusion in ferromagnesian olivine by measuring the O-H distribution in single crystals annealed under hydrothermal conditions. The hydrogen concentration evolves toward an equilibrium value in two stages. First, hydrogen diffuses rapidly (D 11 ≃ 10 -10 m 2/s at 1000 °C) into olivine probably charge compensated by a flux of polarons or of electrons. The resultant metastable equilibrium concentration of hydrogen is directly related to the imposed hydrogen fugacity and to the preexisting concentration of intrinsic point defects; except for electronic species, intrinsic defects diffuse significantly more slowly than hydrogen. Second, hydrogen and intrinsic point defects such as metal vacancies diffuse more slowly (D VMe ≃ 10 -12 m 2/s at 1000 °C) into or out of the olivine as the concentrations of all defect species evolve toward equilibrium with the imposed thermochemical environment.
Article
Taking proper account of the anisotropy factor, gamma , it appears that a single calibration line for the integral molar adsorption coefficient I can be applied as a fair approximation to a variety of substances, including silicate glasses and quartz. This relation is I/gamma = 150(3780 -v) where v is the wavenumber in cm-1 and I is given in cm-2 per mol. H/1. The distribution of hydroxyl bond-strengths is much more biassed to high wavenumbers than at first suggested by the shape of the spectrum.R.A.H.
Article
Several of the supposedly anhydrous major minerals of the upper mantle have been shown to regularly contain small amounts of hydrogen. The concentrations measured in the most important minerals obtained from mantle xenoliths are, expressed in ppm H2O, 100-1300 for clinopyroxene, 60-650 for orthopyroxene, 0-140 for olivine and 1-200 for garnet. Hydrogen is normally structurally incorporated as hydroxyl ions, and in many cases the hydrogen ions seem to act as charge compensators associated with point defects, such as metal vacancies or substitution by mono- or trivalent cations. The determination of the exact amount of hydrogen stored in these nominally anhydrous upper mantle minerals is a key-step toward quantification of the water content of the mantle, as well as understanding of its internal water cycle. For instance, a concentration of 100 ppm H2O homogeneously distributed within the upper mantle above 410 km depth is approximately equivalent to a 100 m water layer at the Earth's surface. However, the relatively fast kinetics of dehydrogenation with concomitant oxidation of iron within these minerals, implies that hydrogen as well as Fe3+ concentrations in equilibrium with mantle conditions might be different from those measured from recovered xenolith samples. High-pressure experimental measurements of hydrogen solubility as a function of p(H(2)O) show a trend similar to the hydrogen contents of natural samples, with hydrogen saturation levels that decrease following the mineral series: diopside > enstatite > olivine > pyrope. Except pyrope, these minerals may incorporate more than 1000 ppm H2O. Based on recent data of water solubility, stability and partitioning, we suggest that an entire upper mantle saturated in hydrogen is highly unprobable and that the maximum average amount of hydrogen stored in the nominally anhydrous minerals of the upper mantle is around 600 ppm H2O. Despite the important progress achieved during the last years, our knowledge of the concentration of hydrogen stored as point defects in the mantle above 410 km is still too poorly constrained. The importance of nominally anhydrous minerals for the water budget of the upper mantle is now well established but still awaits complete quantification.
Article
Computer modeling techniques have been used to study the structure and energetics of H defects in Mg2SiO4 forsterite and their infrared stretching frequencies. The calculated defect binding energies indicate that OH groups will combine with cation vacancies to form neutral defect complexes. Such defect clusters are stabilized by strong hydrogen bonding. Calculated infrared stretching frequencies for OH in different environments have been determined for the first time and show two distinct groups of frequencies. The values of these frequencies, together with their directional components, are compared to experimental measurements. This comparison shows that in pure forsterite, H defects will be associated with both Si and Mg vacancies, but will not occur as isolated OH groups, in line with the results from consideration of defect energetics.