Sie sind auf Seite 1von 12

Tectonophysics 594 (2013) 91102

Contents lists available at SciVerse ScienceDirect

Tectonophysics
journal homepage: www.elsevier.com/locate/tecto

Natural type-C olivine fabrics in garnet peridotites in North Qaidam UHP collision
belt, NW China
Haemyeong Jung a,, Jaeseok Lee a, Byeongkwan Ko a, Sejin Jung a, Munjae Park a, Yi Cao a, Shuguang Song b
a
b

Tectonophysics Laboratory, School of Earth and Environmental Sciences, Seoul National University, Seoul 151-747, Republic of Korea
MOE Key Laboratory of Orogenic Belts and Crustal Evolution, School of Earth and Space Sciences, Peking University, Beijing 100871, China

a r t i c l e

i n f o

Article history:
Received 3 November 2012
Received in revised form 8 March 2013
Accepted 15 March 2013
Available online 24 March 2013
Keywords:
Lattice preferred orientation
Olivine type-C LPO
UHP collision belt
North Qaidam
Water
Seismic anisotropy

a b s t r a c t
Water is known to change the lattice-preferred orientation (LPO) of olivine, which signicantly affects seismic anisotropy in the Earth's upper mantle. Research into the LPO of olivine in the deep interior of the Earth
has been limited due to inadequate specimens. We report both the water-induced LPOs of olivine and the
presence of large quantities of water inside olivine, enstatite, and garnet in garnet peridotites from the
North Qaidam ultrahigh-pressure (UHP) collision belt in NW China. We show that the [001] axis of olivine
is aligned subparallel to the lineation and that the [100] axis is strongly aligned subnormal to the foliation.
This alignment is a known feature of type-C LPO of olivine formed experimentally under water-rich conditions (700 ppm H/Si) at high pressure and temperature. Enstatite possessed an LPO with the [001] axis
aligned parallel to the lineation and the [100] axis aligned normal to the foliation. FTIR analysis of this specimen revealed that olivine contained concentrations of water up to 1130 50 ppm H/Si in clean areas,
whereas olivine, enstatite, and garnet contained considerably more water, i.e., 2600 100 ppm H/Si,
5000 100 ppm H/Si, and 21,000 200 ppm H/Si, respectively, when exsolved inclusions were visible.
Confocal micro-Raman spectroscopy of these exsolved inclusions revealed that they were composed of hornblende and amphiboles. Straight dislocations were also commonly observed in olivine and are characteristic
of olivine that had been experimentally deformed under hydrous conditions. These observations suggest that
the type-C LPO of olivine in the North Qaidam UHP belt formed under water-rich conditions.
2013 Elsevier B.V. Open access under CC BY-NC-ND license.

1. Introduction
Water signicantly affects the physical properties of mantle materials
(Bell and Rossman, 1992; Karato, 2003). Experimental studies conducted
under high pressure and temperature conditions have shown that water
lowers the melting temperature of peridotite (Gaetani and Grove, 1998;
Hirose and Kawamoto, 1995), reduces the ow strength of mantle rocks
(Blacic, 1972; Karato and Jung, 2003; Karato et al., 1986; Mackwell et al.,
1985; Mei and Kohlstedt, 2000), reduces both the bulk and shear moduli
of olivine (Jacobsen et al., 2008), enhances the electrical conductivity of
olivine (Wang et al., 2006; Yoshino et al., 2006), and changes the dominant slip system of olivine (Karato, 1995; Mackwell et al., 1985), which
causes fabric transitions (Jung and Karato, 2001a; Jung et al., 2006;
Katayama et al., 2004). Given that olivine is a dominant mineral in the
upper mantle, a change in its lattice-preferred orientation (LPO) due to
water has signicant implications for interpreting seismic anisotropy in
the upper mantle (Jung, 2009; Jung and Karato, 2001a; Jung et al.,
2006; Karato et al., 2008; Kneller et al., 2005; Long and Silver, 2008,
Corresponding author at: 311 ho, 25-1 dong, Tectonophysics Laboratory, School of
Earth and Environmental Sciences, Seoul National University, Seoul 151-747, Republic
of Korea. Tel.: +82 2 880 6733; fax: +82 2 871 3269.
E-mail address: hjung@snu.ac.kr (H. Jung).
0040-1951 2013 Elsevier B.V. Open access under CC BY-NC-ND license.
http://dx.doi.org/10.1016/j.tecto.2013.03.025

2009; Mainprice, 2007; Mizukami et al., 2004; Nakajima and


Hasegawa, 2004; Park and Levin, 2002; Skemer et al., 2006, 2012;
Tasaka et al., 2008).
Natural rocks, which are rst deformed and then exhumed from
the deep upper mantle, are found in the ultrahigh-pressure (UHP)
belt of several regions and can provide important information on
the deformation conditions and water content deep within the
Earth's interior. Previous studies showed that olivine with a C-type
LPO was observed in samples obtained from the UHP terranes in the
Alpe Arami of Switzerland (Mckel, 1969), Cima di Gagnone in the
Central Alps of Switzerland (Frese et al., 2003), Otroy in Norway
(Katayama et al., 2005; Wang et al., 2013), and Zhimafang garnet peridotites in Eastern China (Xu et al., 2006). However, the water content for olivine in these studies was less than approximately
700 ppm H/Si (Katayama et al., 2005; Xu et al., 2006), which indicates that the olivine LPO might not have formed because of water
(Xu et al., 2006; Wang et al., 2013). In this study, we analyzed the
LPO and water content of olivine within garnet peridotites obtained
from the North Qaidam UHP belt in China. We report the novel discovery of C-type LPOs in olivine containing considerable larger quantities of water than obtained in previous studies, which suggests that
the C-type LPO formed because of water, and a large amount of water
go down subduction zone.

92

H. Jung et al. / Tectonophysics 594 (2013) 91102

2. Geological setting and sample characteristics

The Luliang Shan garnet peridotite massif in North Qaidam in NW


China forms several large lensoid blocks within quartzofeldspathic
gneisses and was exhumed during the Early Paleozoic continental collision between the North China craton and the QaidamQilian block
(Manning et al., 2001; Menold et al., 2001; Song et al., 2003, 2004,
2005a, 2005b, 2006; Yang et al., 2000; Yin et al., 2002, 2007; Zhang et
al., 2005). A geologic map of the garnet peridotite massif is shown in
Fig. 1. The garnet peridotite in the studied area possessed compositional
layering that was largely dened by modal variations in major minerals
such as olivine, orthopyroxene and clinopyroxene (Fig. 2 and gures in
Song et al., 2007). We collected more than ten garnet lherzolites from
the Luliangshan garnet peridotite massif in North Qaidam in NW China
and investigated three of them which showed a well-dened foliation.
The samples consisted of olivine (5068%), enstatite (1133%), diopside
(1017%) and minor garnet. Fig. 3A shows an example of the sample
microstructures in cross-polarized light and demonstrates elongated
clinopyroxene, enstatite, and olivine grains. Exsolution textures were
observed in the garnet, olivine, and enstatite. Exsolutions in garnet consist of rutile, orthopyroxene, clinopyroxene, and amphibole (Fig. 3B, C, D,
and G; see also Song et al., 2004, 2005a, 2009; Yang and Powell, 2008).
Exsolutions in olivine include ilmenite needles, Cr-spinel, and amphibole (Fig. 3E, H, and J). Exsolutions in enstatite (i.e., amphibole) are
also commonly found (Fig. 3F and I). The pyroxene exsolutions found
in garnet (Fig. 3B, C, and D) suggest that the parental garnets originally
possessed excess silicon and were, moreover, majoritic garnets that
are only stable at depths in excess of approximately 150 km (Irifune,
1987; Ringwood and Major, 1971). Given the large amount of pyroxene
exsolutions (up to 4.5 vol.%) and (Fe, Mg)TiO3 exsolution lamellae
(0.530.69 wt.% TiO2) in olivine, the garnet peridotite body in North
Qaidam, China, is believed to have originated at a mantle depth in excess
of 200 km (Song et al., 2004, 2005a, 2007) and belongs to one of the UHP
metamorphic belts, such as the DabieSulu terrane in China (Yang et al.,
1993; Ye et al., 2000; Zhang et al., 2000), Western Gneiss region in
Norway (Van Roermund and Drury, 1998; Van Roermund et al., 2000),
and Alpe Arami in Switzerland (Dobrzhinetskaya et al., 1996; Green
et al., 1997). The application of an Al-in-Opx geothermobarometer
(Brey and Khler, 1990) and a GrtOl geothermobarometer (Oneill

Grt
lherzolite

2 cm
North
Qaidam,
China

Fig. 2. (A) An example of foliated garnet lherzolite in North Qaidam. (B) Compositional
layering of Grtlherzolite and a pyroxene-rich layer. The dark color shows weathered
olivine in the surface. Length of the pen is 14 cm. (C) Compositional layering of Grt
lherzolite and Grtdunite. Purple grains represent fresh garnets. Grt: garnet, Py:
pyroxene.

and Wood, 1979) to the matrix host minerals within the garnet peridotite yielded a pressure of P = 5.06.5 GPa and a temperature of
T = 9601040 C (Song et al., 2004). Olivine shows a wide range of
Fo values (Mg#, Mg/[Mg + Fe 2+] = 0.830.91) (Song et al., 2007).
Further details on the chemical composition of olivine, orthopyroxene,
and garnet in our samples are available from previous studies (Song et
al., 2004, 2007).
3. Methods
3.1. Determination of the LPO in minerals and seismic
anisotropy calculations
Fig. 1. Geologic map of the Luliangshan garnet peridotite massif in North Qaidam, NW
China.
Modied after Song et al., 2007.

The foliation of each sample was determined by the compositional


layering of olivine, pyroxene, and garnet, and the lineation of each

H. Jung et al. / Tectonophysics 594 (2013) 91102

93

Fig. 3. (A) Optical photomicrograph of sample 525 using cross-polarized light. Grt: garnet, Ol: olivine, En: enstatite, and Cpx: clinopyroxene. Sample 525 shows a porphyroclastic
texture and both En and Cpx are elongated subparallel to the lineation (EW direction). (B) An optical photomicrograph showing an enstatite inclusion in garnet. (C) Backscattered
electron (BSE) image showing En and Cpx in garnet. (D) BSE image showing enstatite and amphibole (tremolite) in garnet. (E) Optical photomicrograph of sample 518 using
cross-polarized light, which shows Cr-spinel and amphibole in olivine. (F) Optical photomicrograph using transmitted plane-polarized light and showing amphibole lamellae in
enstatite (sample 518). (G) BSE image showing hornblende in garnet. (H) Optical photomicrograph using transmitted plane-polarized light, which shows ilmenite rods and amphibole lamellae (sample 525). (I) Optical photomicrograph using transmitted plane-polarized light and showing amphibole lamellae in enstatite (sample 525). (J) BSE image
showing amphibole (anthophyllite) in olivine (sample 525).

sample was determined from the grain elongation determined by grain


shape analysis of digitized lines in the foliation (Panozzo, 1983, 1984).
The samples were cut parallel to their lineation and perpendicular to
their foliation for the microstructure analysis. The LPO of olivine and
enstatite was measured for each sample by electron backscatter diffraction (EBSD) (Jung et al., 2006; Prior et al., 1999) using a scanning electron
microscope (SEM JEOL 6380) at the School of Earth and Environmental
Sciences (SEES), Seoul National University (SNU) in Korea, and the
EBSD patterns were analyzed using HKL Channel 5 software and manually indexed to ensure the solution accuracy. The SEM operated at an
accelerating voltage of 20 kV and a working distance of 15 mm. Approximately 150250 individual olivine and enstatite grains were measured for each sample. The misorientation index (M-index) (Skemer et
al., 2005) was calculated to estimate the fabric strength of the sample
using the uncorrelated grain pairs determined from the EBSD data.
The seismic velocity and seismic anisotropy were calculated from the
orientation data for olivine, enstatite, diopside, and garnet using the

elastic constants for olivine (Abramson et al., 1997), enstatite (Chai et


al., 1997), diopside (Collins and Brown, 1998), garnet (Bass, 1989),
and a software program (Mainprice, 1990).
3.2. Measurement of the mineral water content
The water content of olivine, enstatite, and garnet was measured
using a Nicolet 6700 FTIR spectrometer with a Continuum IR microscope housed at the Tectonophysics Laboratory at the SEES in SNU,
Korea. The FTIR spectrometer has a spectral range of 7800350 cm1,
electronically temperature controlled Ever-Glo source, DLaTGS w/KBR
Detector and Ge on KBR beamsplitter with OMNIC software. Each sample was sliced to a thickness of 300 m, polished on both sides to a
thickness of 210 m, and dried at 120 C for over 24 h prior to the
FTIR analysis. Next, the sample chamber was purged with N2 gas during
the FTIR measurements to exclude atmospheric moisture from our
measurements. An IR beam with the dimensions of 50 50 m was

94

H. Jung et al. / Tectonophysics 594 (2013) 91102

focused on a single crystal while avoiding cracks and grain boundaries using unpolarized transmitted light. The number of scans accumulated to obtain the FTIR spectrum was 128. The Paterson calibration
(Paterson, 1982) was used to calculate the amount of water from infrared absorptions in the range of 30003750 cm1, which is the region
dominated by the O\H stretching vibrations. The background water
content was determined in air immediately before measuring the specimen. Since the original paper discussing the effects of water on the
LPO of olivine (Jung and Karato, 2001a) used the Paterson calibration
and unpolarized light to determine the water content, we decided to
report the mineral water contents using the Paterson calibration and
unpolarized light.
3.3. Observation of dislocations
The samples were decorated with oxygen (Kohlstedt et al., 1976) for
1 h at 800 C and polished using Syton (colloidal silica) to remove the
thin oxide layer that formed on their surface. After this oxygen decoration, the dislocations in olivine were imaged using a scanning electron
microscope (SEM JEOL 6380) (Jung and Karato, 2001b; Karato, 1987)
at the SEES in SNU, Korea. A back-scattered electron (BSE) image was
taken using a 15 kV acceleration voltage, 20 nA beam current, and 60
spot size.
3.4. Micro-Raman spectroscopy
A dispersive confocal DXR Raman microscope (Thermo Scientic)
housed at the Tectonophysics Laboratory of the SEES in SNU, Korea
was used to identify exsolved lamellae in garnet, olivine, and enstatite.
The Raman microscope was equipped with a 532 nm laser (10 mW
power) and an optical microscope (Olympus, 50 objective) with an automatic stage and had a resolution of 0.01 cm1 over the wavenumber
range of 503550 cm 1 with a beam size of 0.67 m, which is small
enough to analyze small inclusions. The Raman spectrum was mostly
obtained at a depth of 1050 m below the specimen surface using a
32-s exposure time.
4. Results
4.1. LPO of olivine and enstatite
The LPOs for both olivine and enstatite are shown in the pole
gure (Fig. 4). Samples 518 and 525 showed that the [001] axis of
olivine was aligned subparallel to the lineation, whereas the [100]
axis was strongly aligned subnormal to the foliation; this alignment
is characteristic of the C-type LPO in olivine (Jung and Karato,
2001a; Jung et al., 2006). Sample 524 showed similar C-type LPO patterns for olivine in both the [100] and [001] axes; however, the [010]
axis had a tendency to align subparallel to the lineation. There was an
insufcient amount of EBSD data for the enstatites in samples 518
and 524 (see Supplementary Fig. S1), and the LPO of enstatite was
only shown for sample 525 (Fig. 4D). The [001] axis of enstatite was
strongly aligned parallel to the lineation, and the [100] axis was aligned
subnormal to the foliation; such alignment is characteristic of the
AC-type LPO of enstatite (Jung et al., 2010). LPOs of other minor minerals such as diopside and garnet are shown in Supplementary Fig. S1.
Combined pole gure of each mineral for all three samples is shown
in Supplementary Fig. S3. It was also observed by EBSD that the ilmenite
rods in olivine were aligned parallel to the [010] axis of olivine.

in olivine (Fig. 5CE); 3689, 3682, 3676, 3649, 3630, 3618, 3588, 3567,
3529, 3524, 3516, and 3424 cm 1 in enstatite (Fig. 5A & F); and 3687,
3675, 3670, 3649, 3629, 3620, 3609, 3587, 3567, and 3546 cm 1 in garnet (Fig. 5B). Other samples showed similar IR peaks. The IR peaks at approximately 3689, 3687, and 3685 cm1 can be attributed to serpentine
(Jung, 2009; Khisina et al., 2001; Mellini et al., 2002; Miller et al., 1987;
Post and Borer, 2000); the peaks at approximately 3676 and 3660 cm1
can be associated to amphibole (Skogby and Rossman, 1991) or talc
(Khisina et al., 2001); while those at approximately 3613, 3598, 3571,
3567, and 3525 cm1 represent structurally bound water in olivine
(Kohlstedt et al., 1996; Mei and Kohlstedt, 2000). Olivine, enstatite, and
garnet contained a large amount of water, i.e., 2600 100 ppm H/Si,
5000 100 ppm H/Si, and 21,000 200 ppm H/Si, respectively; however, it should be noted that exsolved hydrous inclusions were present in
olivine, enstatite, and garnet (Fig. 5AC). The obtained values indicate an
average water content of 1014 grains. We also analyzed the water content of olivine without visible hydrous inclusions under transmitted light
in an optical microscope. The water content in the clean olivine was
930 50 ppm H/Si in average for 11 olivine grains. We didn't observe
any difference in water content between large and small olivine. A
water content of 1130 50 ppm H/Si in clean olivine in this study is
the largest amount of water found in natural olivine with type-C LPO in
the world. A sample FTIR spectrum of the clean olivine is shown in
Fig. 5E; the water content in this olivine was 1130 50 ppm H/Si. The
varying water content of clean olivines indicates hydrogen diffused
from the olivine during uplift. The water content reported here would
be approximately 3 times more than that estimated in this study using
the alternate calibration (Bell et al., 2003).
4.3. Micro-Raman analysis of hydrous inclusions
Several hydrous inclusions were found in olivine, enstatite, and garnet. These inclusions were analyzed using a confocal DXR micro-Raman
microscope. The micro-Raman analysis of specimen (525) revealed that
the exsolved olivine, enstatite, and garnet inclusions were amphiboles
(Fig. 6). Hornblende and tremolite were both found in garnet (Fig. 6A),
crocidolite was observed in enstatite (Fig. 6B), and anthophyllite was
found in olivine (Fig. 6C, D). A Raman shift at 673 cm1 for the exsolved
inclusion in garnet (Fig. 6A: Tre) indicates the presence of tremolite
(Bard et al., 1997; Blaha and Rosasco, 1978; Lewis et al., 1996). The
Raman shifts at 664, 581, and 248 cm1 appeared in the exsolved garnet inclusions (Fig. 6A: Hb) and indicate the presence of hornblende,
which was conrmed by the chemical composition of the mineral determined using a JEOL JXA-8900R electron probe microanalyzer (EPMA) at
SNU, Korea. The Raman shifts at 581 and 775 cm1 for the exsolved inclusion in enstatite (Fig. 6B: Cro) indicate the presence of crocidolite
(Bard et al., 1997; Lewis et al., 1996; Petry et al., 2006). The Raman shifts
at 675, 670, 538, 429, 337, and 190 cm 1 for the exsolved olivine inclusions (Fig. 6C, D: Anth) indicate the presence of anthophyllite (Bard et
al., 1997; Rinaudo et al., 2004).
4.4. Dislocation microstructures of olivine
Dislocation microstructures were observed in olivine using backscattered electron (BSE) images (Fig. 7) in a scanning electron microscope after the oxygen decoration of the specimens. These dislocations
are primarily straight in olivine, which is consistent with the observation of straight dislocations in olivine with C-type LPO in previous experimental study at high pressure and temperature (Jung and Karato,
2001a).

4.2. Water content of olivine, enstatite, and garnet


4.5. Stress estimation of sample
We observed a large amount of water in olivine, enstatite, and garnet, and representative FTIR spectra (sample 525) are shown in Fig. 5.
IR peaks were observed at the following wave numbers: 3689, 3676,
3668, 3649, 3630, 3613, 3598, 3588, 3568, 3564, 3536, and 3522 cm1

Further deformation conditions associated with the formation of


C-type LPO in olivine from North Qaidam were inferred from the following microstructural observations. A recrystallized grain-size piezometer

H. Jung et al. / Tectonophysics 594 (2013) 91102

[100]

A. 518
N = 155
M = 0.09

95

[010]

[001]

L 1
2
3
4
5

B. 524
N = 195
M = 0.08

1
2
3

C. 525
N = 248
M = 0.06

1
2
3

D. 525 En
N = 167
M = 0.22

1
3
5
7

Fig. 4. Pole gures of olivine (AC) and enstatite (D) presented in the upper hemisphere and equal area projection. L and S represent the lineation and foliation, respectively. Olivine
shows a typical C-type LPO. The color coding represents the data point density. The contours correspond to multiples of the uniform distribution. A half width of 20 was used to
draw the pole gures. N: number of grains analyzed. M: Misorientation-index (Skemer et al., 2005) to show the fabric strength. En: enstatite.

can be used to estimate the differential stress of a sample (Jung and


Karato, 2001b; Karato et al., 1980; van der Wal et al., 1993). Average
recrystallized grain-sizes in olivine samples 518 and 525 of 435 m
and 530 m, respectively, were measured using the linear intercept
method (Gifkins, 1970; Russ and Dehoff, 2000) and indicate differential
stresses of 41 15 MPa and 36 MPa 15 MPa, respectively, for the
samples under wet conditions (Jung and Karato, 2001b). Moreover,
the dislocation density () of olivine can be used to independently estimate the differential stress of each sample (Jung and Karato, 2001b;
Karato and Jung, 2003), and can also be estimated by dividing the
total dislocation length by the total grain volume (Jung and Karato,
2001b). The distribution of dislocation densities for 30 grains (sample
518) is shown in Fig. 8 and was determined from backscattered electron
images. The mean density was calculated to be = 3.25 10 11/m2
and = 2.19 1011/m 2 for samples 518 and 525, respectively. Furthermore, the stresses ( = 55 20 MPa and = 40 20 MPa)
from the dislocation density in samples 518 and 525 were calculated
using the relationship between the stress and the dislocation density
(Jung and Karato, 2001b). The differential stress for two samples was
estimated from the dislocation density and was in good agreement
with those estimated from the recrystallized grain-size piezometer.
4.6. Seismic anisotropy
The seismic P- and S-wave anisotropy (Vp and AVs) (Fig. 9) was calculated using the LPO of olivine and enstatite corresponding to the

fabrics (Fig. 4). The P-wave (Vp) exhibited seismic anisotropy of olivine
in the range of 2.75.9% for three samples. The Vp anisotropy of
enstatite (sample 525) was 3.4%. The maximum anisotropy of the
shear wave (AVs) in olivine was in the range of 2.445.02%. By combining the seismic anisotropy of both enstatite and olivine (sample 525),
we found that both the P-wave and the S-wave anisotropies were reduced to 1.5% and 2.3%, respectively (Fig. 9E), which indicates that the
LPO of enstatite reduced the overall seismic anisotropy of the sample
(e.g., Mainprice and Silver, 1993). The polarization direction of the fast
S-wave (VS1) in the two samples (518, 524) was normal to their lineation (EW) as shown in Fig. 9A and B, whereas that of sample 525
was subnormal to the lineation. By combining the polarization anisotropy of enstatite (Fig. 9D) with that of olivine for sample 525, we found
that the polarization direction of the fast S-wave shifted nearly 45
from the north at the center of the pole gure (Fig. 9E), which depicts
the vertically propagating S-waves. This result indicates that the LPO
of enstatite can modify the overall polarization direction of the fast
S-waves. In addition, we calculated seismic properties of each sample
considering all constituent minerals (olivine, enstatite, diopside, and
grossular) (Supplementary Fig. S2). Seismic anisotropy of combined
LPOs (Fig. S3) of each mineral for all three samples (518, 524, and
525) is shown in Supplementary Fig. S4. Elastic stiffness (Cij) of Vogit
ReussHill average was calculated using the LPO and density of mineral.
Elastic stiffness (Cij) of representative olivine (518) and enstatite (525)
is shown in Table 1. Elastic stiffnesses (Cij) of combined LPOs of individual mineral for all three samples are shown in Supplementary Table S1.

96

H. Jung et al. / Tectonophysics 594 (2013) 91102

3567

3527

3649
3629
3608

3689
3676
3668

C. OI

Fig. 5. Representative unpolarized FTIR spectra of (A and F) enstatite, (B) garnet, and (CE) olivine. The inset images show a 50 m 50 m square, which is the area of the transmitted IR beam. (AC) FTIR spectra of sample 525, which contains hydrous exsolutions such as amphibole and serpentine. The water content of mineral based on spectra A, B, C and
D was 9000 200 ppm H/Si, 11900 200 ppm H/Si, 2600 100 ppm H/Si and 1120 50 ppm H/Si, respectively. (D) Many ilmenite rods are visible in the olivine. (EF) FTIR
spectra of sample 525. The IR beam passed through a clean area of olivine (E) and enstatite (F). The water content in this clean area (E) was 1130 50 ppm H/Si for olivine.

5. Discussion and implications


5.1. LPO of olivine
We have reported the presence of C-type LPO in olivine within
garnet lherzolites obtained from the North Qaidam UHP belt for the
rst time. Similar C-type LPO was found in olivine from other UHP
terranes around the world, such as the Alpe Arami in Switzerland
(Mckel, 1969), Cima di Gagnone in the Central Alps in Switzerland
(Frese et al., 2003), Otroy in Norway (Katayama et al., 2005; Wang
et al., 2013), and Zhimafang garnet peridotite in the Sulu UHP terrane
in China (Xu et al., 2006). Because samples with a C-type LPO did not
contain a considerable amount of water, Xu et al. (2006) proposed
that the fabric was formed by high pressure and low temperature.
In this case, water may have been diffused out of olivine after forming
the C-type LPO because of the high hydrogen diffusivity of olivine
(Kohlstedt and Mackwell, 1998). The C-type LPO of olivine was experimentally produced under low stress and water-rich conditions
(700 ppm H/Si) (Jung and Karato, 2001a; Jung et al., 2006). In this
study, we report olivine showing C-type LPO and containing a large

amount of water, i.e., 1130 50 ppm H/Si, for the rst time when
hydrous inclusions in olivine were not visible (Fig. 5E). Both this observation and a previous experimental study (Jung and Karato,
2001a) suggest that C-type LPO formed in olivine because of water.
5.2. Effect of pressure on the LPO of olivine
It has been suggested that pressure may affect the LPO of olivine
(Couvy et al., 2004; Jung, 2012; Jung et al., 2009; Mainprice et al., 2005;
Ohuchi et al., 2011). Couvy et al. (2004) performed relaxation experiments on Fe-free olivine (forsterite) at a pressure of 11 GPa and temperature of 1400 C and reported that a C-type LPO formed due to the high
pressure. As previously suggested (Jung et al., 2009; Karato et al., 2008;
Katayama et al., 2005), these samples (Couvy et al., 2004) contained a
large amount of water (approximately 8502200 ppm H/Si) and experienced a differential stress of approximately 1001500 MPa, which can
induce [001] slipping in olivine (Carter and Av Lallemant, 1970; Jung
and Karato, 2001a). Therefore, the C-type LPO produced by Couvy et al.
(2004) may have been produced by either water or high stress.
Deforming a single crystal of Fe-free olivine (forsterite), Raterron et al.

680

436

535

746
681

1008

398

660

207

357

547

673

522
436

581
538

775

860

942

166
122

224

248

357

423

515

554

664

Tre
581

915
861

916

338

En
Grt

1020

860

557
638

859

1046

938

357

396

659

917

1007

97
338

H. Jung et al. / Tectonophysics 594 (2013) 91102

Cro

337
298

424

538

681
660
585

1007
300

424

587
539

715

957
917

954
916

853

852
820

822

Hb

Ol

107

190
289
232

337

359
465
429

576
538

791

852
821

1007

Anth

953
917

298

424

538

523

670

959
915

852
820

675

Ol

Anth

Fig. 6. Raman spectra of both the host mineral and the exsolved hydrous mineral within it. (A) Raman spectra of garnet and its exsolved minerals. Grt: host garnet, Tre: exsolved
amphibole (tremolite) in garnet, and Hb: exsolved hornblende in garnet. (B) Raman spectra of enstatite and exsolved inclusion. En: host enstatite, Cro: exsolved amphibole
(crocidolite). (C) Raman spectra of olivine and exsolved amphibole. Ol: host olivine, Anth: exsolved amphibole (anthophyllite) in olivine. (D) Raman spectra of another olivine
and exsolved amphibole. Ol: host olivine, Anth: exsolved amphibole (anthophyllite) in olivine. Grt: garnet, Tre: tremolite, Hb: hornblende, En: enstatite, Cro: crocidolite, Ol: olivine,
and Anth: anthophyllite.

(2007) showed that the [001] slip was easier than the [100] slip at pressures between 3 and 7 GPa. Moreover, recent experimental study of
harzburgite (Jung et al., 2009) showed that olivine deformed at low
stress and dry condition under pressures higher than 3 GPa did not produce C-type LPO, but B-type LPO. In addition, Ohuchi et al. (2011) also
produced B-type LPO of olivine at high pressure of 7 GPa. These studies
indicate that the C-type LPO of olivine was not produced by high pressure under dry conditions. To the best of our knowledge, there is no
solid experimental evidence that C-type LPO in olivine forms from
high pressure alone. In addition, natural C-type LPOs of olivine were
found recently in the spinel peridotites which were originated from a
depth of 4060 km at Rio Grande rift, U.S.A. (Park et al., submitted for
publication), suggesting that C-type LPO of olivine can be also formed
at low pressure condition and high pressure is not a prerequisite to
make C-type LPO.
5.3. Origin of garnet peridotite in the North Qaidam UHP belt
The origin of garnet peridotite in North Qaidam is still controversial. Yang and Powell (2008) argued that the Luliangshan garnet peridotites probably formed via the dehydration of oceanic serpentinized
peridotites during the subduction of the oceanic slab. Based on ReOs
data and bulk-rock geochemistry, Shi et al. (2010) suggested that

the Shenglikou (Luliangshan) dunites are derived from the Archean


(>2.6 Ga) subcontinental lithospheric mantle (SCLM), and garnet
lherzolites were produced through the refertilization of these dunites
by mac melts (now pyroxenites) during the Paleoproterozoic era (ca.
2.2 Ga). Based on zircon Hf isotopic data, Xiong et al. (2011) recently
suggested that the Shenglikou (Luliangshan) garnet peridotite massif
is a fragment of metasomatized Archean SCLM from the Qilian block.
However, petrological and geochemical data do not seem to support
the peridotite massif being a suboceanic or subcontinental lithospheric
mantle fragment. The large compositional variation of the primary
rock-forming minerals, the documented major and trace element systematics, and the magmatic zircon cores strongly suggest that the
Luliangshan garnet peridotite ultramac complex crystallized from
high-Mg melts in an arc environment before UHP metamorphism
(Song et al., 2007). It was previously suggested that corner ow of the
mantle wedge during continental subduction lowered the garnet peridotites to depths of approximately 150200 km after approximately
420 Ma and exhumed the garnet peridotites during subsequent continental collision (approximately 400 Ma) (Song et al., 2005b; Song et
al., 2007). Based on the study of thermomechanical modeling, Yin et
al. (2007) suggested that the development of UHP metamorphism and
subsequent exhumation of the UHP metamorphic rocks in the North
Qaidam area are related to diapiric ows during oceanic subduction.

98

H. Jung et al. / Tectonophysics 594 (2013) 91102

12

10

Frequency

0.8

1.8

2.8

3.8

4.8

5.8

Dislocation density (x 1011 m-2 )


Fig. 8. Dislocation density distribution for 30 olivine grains in sample 525. The average
dislocation density of olivine in the sample was = 2.19 1011/m2.

Fig. 7. Backscattered electron (BSE) images showing typical dislocation microstructures within olivine (sample 525). Straight dislocations were commonly observed.
The white lines and dots denote dislocations.

5.4. Stress estimation of sample


Stress of specimen was estimated by two methods in this study:
recrystallized grain-size piezometer and dislocation density. Stress
estimation from dislocation density needs to be cautious because dislocation density of olivine can be affected by annealing or later deformation processes. In our study, stress estimated from recrystallized
grain-size of olivine was similar to the one estimated from dislocation
density, indicating that dislocation density of specimen was not much
affected by later deformation processes.
5.5. Evidence of large amount of water in peridotites
A large amount of water was found to be present in olivine in garnet
lherzolite from the orogenic UHP belt in North Qaidam. We detected
water contents of up to 1130 50 ppm H/Si in clean olivine with an
average value of 2600 100 ppm H/Si in areas containing exsolved
amphibole. Experiments on the solubility of water in olivine (Jung
et al., 2006; Kohlstedt et al., 1996) indicate that it increases with increasing pressure. In our study, the water content of olivine (1130
2600 100 ppm H/Si) corresponds to the solubility of water in olivine
at a pressure of 23 GPa and temperature of 1300 C. This is considered
to be the minimum water content in olivine because hydrogen in olivine
might have been diffused out during the exhumation of rocks to surface
because of high diffusivity of hydrogen in olivine (Kohlstedt and
Mackwell, 1998). Furthermore, the average water content detected in
garnet was approximately 21,000 200 ppm H/Si. Previous experimental studies on the solubility of water in garnet (Lu and Keppler,
1997; Withers et al., 1998) showed that it also increased with increasing

pressure. The water content of garnet (21,000 ppm H/Si, which is


equivalent to 1460 ppm H2O) corresponds to the solubility of water
above approximately 5 GPa (Withers et al., 1998). We found amphibole
lamellae in olivine, enstatite, and garnet (Figs. 3, 5, and 6). Sodic amphiboles in garnet from the same locality were reported previously (Song
et al., 2005a, 2005b, 2007; Yang and Powell, 2008), and amphibole lamellae in olivine have also been reported (Yang and Powell, 2008).
Song et al. (2005a, 2005b) found a topotaxial relationship between
the lamellae and host garnet via TEM observation, which suggests that
these amphibole lamellae formed during decompression-induced exsolution from a parental garnet. The presence of amphiboles in olivine,
enstatite, and garnet implied that the olivine, enstatite, and garnet precursors were OH-bearing minerals that were stable at the mantle depth
during a continental collision in the early Paleozoic (approximately
423 Ma by SHRIMP dating of zircon) (Song et al., 2005b). The results
from our FTIR study indicated that olivine, enstatite, and garnet from
orogenic garnet peridotite in the cold subduction zones may contain a
considerably larger number of hydroxyls than those obtained from orogenic garnet peridotite in the hot subcontinental mantle (Dixon et al.,
2002).

5.6. Possible origin of water in the specimen


Song et al. (2007, 2009) suggested three stages of evolution for
Luliangshan garnet peridotite. Stage I: cooling of Mg-rich melts led
to the formation of cumulate ultramac assemblage in a sub-arc lithospheric mantle overlaying a paleo-subduction zone; Stage II: oceanic
lithosphere subduction induced asthenospheric corner ow which
transported the cumulate peridotite body deep into the mantle in
the subduction zone; and Stage III: the subducted continental crust
seized the peridotite body and carried it to depths in excess of
200 km. It is unlikely that large quantities of water dissolved in its
original cumulative rocks. Therefore, the source of water preserved
in the anhydrous minerals was probably originated from the cold
and hydrous subducted oceanic lithosphere at Stage II. The olivine
C-type LPO and estimated stress of 4050 MPa suggest that the deformation temperature is supposed to be higher than 750850 C
(Katayama and Karato, 2006). The calculated PT conditions of garnet
lherzolite (P = 5.06.5 GPa, T = 9601040 C) and garnet-bearing
dunite (P = 4.65.3 GPa, T = 9801130 C) (Song et al., 2009) are
consistent with required temperature to form C-type LPO. In addition,

H. Jung et al. / Tectonophysics 594 (2013) 91102

Vp Contours (km/s)

AVs Contours (%)


8.61

(A) 518

99

Vs1 Polarization
5.03

5.03

8.55
4.0

Ol

8.45

3.0

8.35

2.0

8.25

1.0

8.17

0.14

0.14

8.65

4.35

4.35

Anisotropy = 5.1 %

(B) 524

8.55
3.0

Ol
8.40
8.25

2.0
1.0

8.15

0.12

0.12

8.47

2.45

2.45

Anisotropy = 5.9 %

(C) 525

2.00
8.40

Ol

1.50
8.35
1.00
8.30
.50

8.24

0.08

0.08

8.26

3.87

3.87

Anisotropy = 2.7 %

(D) 525
8.20

3.0

En
2.0
8.10
8.05

1.0

7.99

0.10

0.10

8.32

2.33

2.33

Anisotropy = 3.4 %

(E) 525

8.30

Ol+En

2.00
1.50

8.26

8.22

8.20

1.00
.50

0.02

0.02

Anisotropy = 1.5 %
Fig. 9. Seismic anisotropy calculated for the fabrics shown in Fig. 4. Seismic anisotropies are shown in the upper hemisphere and equal area projection. The EW directions correspond to the lineation (L), and the horizontal line represents the foliation. Both the compressional-wave velocity (Vp) and shear-wave anisotropy (AVs) are shown. Vs1 is a plot of
the polarization direction of fast S-waves along different propagation orientations. The center of the stereonet corresponds to a vertical propagation. For the vertical propagation of
S-waves in olivine (AC), the polarization direction is nearly perpendicular to the ow direction (lineation). Small black square and white circle represent maximum and minimum
anisotropy, respectively. Ol: olivine, En: enstatite.

this minimum temperatures (750850 C) forming C-type LPO is


more or less coincident with the initial temperature (~ 700800 C)
(corresponding to ~3 GPa) of exhumation of UHP metamorphic rocks
(e.g., Yin et al., 2007). Therefore, we can infer that the deformation

probably occurred during the subduction of continental crust above


3 GPa (Stage III). Because the solubility of water in olivine increases
with increasing pressure (Kohlstedt et al., 1996), the water uptaken
in Stage II can be well preserved at high pressure and secure the

100

H. Jung et al. / Tectonophysics 594 (2013) 91102

Table 1
Elastic stiffness, Cij (GPa) of representative olivine (sample 518) and enstatite (sample
525) at ambient condition. Reference axes dened as 2: lineation, 1: normal to foliation,
and 3: perpendicular to both 1 and 2 directions (e.g., Z = 1, X = 2, Y = 3).
i\j

Olivine
1
2
3
4
5
6
Density

Sample 518
247.87

Enstatite
1
2
3
4
5
6
Density

76.37
225.47

77.49
77.44
229.84

0.2
0.85
0.58
74.24

0.75
0.27
1.45
0.68
81.1

1.77
1.22
0.76
0.52
1.03
77.76

61.9
224.71

68.61
59.9
213.71

1.56
1.8
0.89
79.39

2.4
1.08
2.64
0.15
75.14

1.12
0.44
2.04
0.57
1.11
79.66

3.355 g/cm3
Sample 525
220.58

3.306 g/cm3

deformation to occur in a wet condition. However, origin of water in


the peridotites in North Qaidam UHP belt is not yet clear. We need a
further study to understand that where, when, and how rocks became hydrated.
6. Conclusions
The C-type LPO of olivine a fabric reecting the (100)[001] slip
system was found in garnet lherzolites obtained from the North
Qaidam UHP collision belt. An FTIR study of the samples yielded
water contents for olivine, enstatite, and garnet of 2600, 5000, and
21,000 200 ppm H/Si, respectively, when exsolved hydrous inclusions were found inside the crystals. Micro-Raman analysis of
specimens 518 and 525 revealed these exsolved olivine, enstatite,
and garnet inclusions to be amphiboles from the decompressioninduced exsolution of a parental olivine, enstatite, and garnet. The
water content in olivine was as high as 1130 50 ppm H/Si when
hydrous inclusions in olivine were not visible. This is the largest
amount of water found thus far in natural olivine showing a C-type
LPO and exceeds the minimum water content (700 ppm H/Si) required for the C-type LPO to be produced in olivine during high pressure experiments (Jung and Karato, 2001a), which indicates that the
type-C LPO of olivine was produced due to water. The sample stresses were determined from the grain-size piezometer and the dislocation density of olivine and turned out to be 3555 20 MPa. These
results are consistent with those of previous experimental studies
(Jung and Karato, 2001a; Jung et al., 2006), which found the C-type
LPO of olivine formed under low stress and wet conditions (water
content 700 ppm H/Si). Our results showed that rocks showing the
type-C LPOs of olivine have lots of water and they were originated
from a UHP collision zone, implying that large amount of water go
down subduction zones and induce changes in olivine fabric and seismic anisotropy.
Supplementary data to this article can be found online at http://
dx.doi.org/10.1016/j.tecto.2013.03.025.
Acknowledgments
This work was supported by the Mid-career Research Program
through an NRF grant funded by the MEST (3345-20100013) in Korea
and National Natural Science Foundation of China (Nos. 40825007,
41121062). H. J. thanks S. Ryu for her help in FTIR study at the initial
stage of project. We thank A. Yin and anonymous reviewers for their
constructive comments.

References
Abramson, E.H., Brown, J.M., Slutsky, L.J., Zaug, J., 1997. The elastic constants of San Carlos
olivine to 17 GPa. Journal of Geophysical ResearchSolid Earth 102, 1225312263.
Bard, D., Yarwood, J., Tylee, B., 1997. Asbestos bre identication by Raman
microspectroscopy. Journal of Raman Spectroscopy 28, 803809.
Bass, J.D., 1989. Elasticity of grossular and spessartite garnets by Brillouin spectroscopy.
Journal of Geophysical Research 94, 76217628.
Bell, D.R., Rossman, G.R., 1992. Water in Earth's mantle: the role of nominally anhydrous minerals. Science 255, 13911397.
Bell, D.R., Rossman, G.R., Maldener, J., Endisch, D., Rauch, F., 2003. Hydroxide in olivine:
a quantitative determination of the absolute amount and calibration of the IR spectrum. Journal of Geophysical ResearchSolid Earth 108, 2105.
Blacic, J.D., 1972. Effects of water in the experimental deformation of olivine. In: Heard,
H.C., Borg, I.Y., Carter, N.L., Raleigh, C.B. (Eds.), Flow and Fracture of Rocks. American
Geophysical Union, Washington DC, pp. 109115.
Blaha, J.J., Rosasco, G.J., 1978. Raman microprobe spectra of individual microcrystals
and bers of talc, tremolite, and related silicate minerals. Analytical Chemistry
50, 892896.
Brey, G.P., Khler, T., 1990. Geothermobarometry in four-phase lherzolites II. New
thermobarometers, and practical assessment of existing thermobarometers. Journal
of Petrology 31, 13531378.
Carter, N.L., Av Lallemant, H.G., 1970. High temperature ow of dunite and peridotite.
Bulletin of the Geological Society of America 81, 21812202.
Chai, M., Brown, J.M., Slutsky, L.J., 1997. The elastic constants of an aluminous
orthopyroxene to 12.5 GPa. Journal of Geophysical ResearchSolid Earth 102,
1477914785.
Collins, M.D., Brown, J.M., 1998. Elasticity of an upper mantle clinopyroxene. Physics
and Chemistry of Minerals 26, 713.
Couvy, H., Frost, D.J., Heidelbach, F., Nyilas, K., Ungar, T., Mackwell, S., Cordier, P., 2004.
Shear deformation experiments of forsterite at 11 GPa1400 degrees C in the
multianvil apparatus. European Journal of Mineralogy 16, 877889.
Dixon, J.E., Leist, L., Langmuir, C., Schilling, J.G., 2002. Recycled dehydrated lithosphere
observed in plume-inuenced mid-ocean-ridge basalt. Nature 420, 385389.
Dobrzhinetskaya, L., Green, H.W., Wang, S., 1996. Alpe Arami: a peridotite massif from
depths of more than 300 kilometers. Science 271, 18411845.
Frese, K., Trommsdorff, V., Kunze, K., 2003. Olivine [100] normal to foliation: lattice
preferred orientation in prograde garnet peridotite formed at high H2O activity,
Cima di Gagnone (Central Alps). Contributions to Mineralogy and Petrology 145,
7586.
Gaetani, G.A., Grove, T.L., 1998. The inuence of water on melting of mantle peridotite.
Contributions to Mineralogy and Petrology 131, 323346.
Gifkins, R.C., 1970. Optical Microscopy of Metals. Elsevier, New York.
Green, H.W., Dobrzhinetskaya, L., Bozhilov, K., 1997. Determining the origin of
ultrahigh-pressure lherzolitesresponse. Science 278, 704707.
Hirose, K., Kawamoto, T., 1995. Hydrous partial melting of lherzolite at 1 GPathe
effect of H2O on the genesis of basaltic magmas. Earth and Planetary Science
Letters 133, 463473.
Irifune, T., 1987. An experimental investigation of the pyroxene garnet transformation
in a pyrolite composition and its bearing on the constitution of the mantle. Physics
of the Earth and Planetary Interiors 45, 324336.
Jacobsen, S.D., Jiang, F.M., Mao, Z., Duffy, T.S., Smyth, J.R., Holl, C.M., Frost, D.J., 2008.
Effects of hydration on the elastic properties of olivine. Geophysical Research
Letters 35. http://dx.doi.org/10.1029/2008gl034398.
Jung, H., 2009. Deformation fabrics of olivine in Val Malenco peridotite found in Italy
and implications for the seismic anisotropy in the upper mantle. Lithos 109,
341349.
Jung, H., 2012. Rock deformation and formation of LPO of minerals in the upper mantle:
implications for seismic anisotropy. Journal of the Petrological Society of Korea 21,
249261.
Jung, H., Karato, S., 2001a. Water-induced fabric transitions in olivine. Science 293,
14601463.
Jung, H., Karato, S.I., 2001b. Effects of water on dynamically recrystallized grain-size of
olivine. Journal of Structural Geology 23, 13371344.
Jung, H., Katayama, I., Jiang, Z., Hiraga, T., Karato, S., 2006. Effect of water and stress on
the lattice-preferred orientation of olivine. Tectonophysics 421, 122.
Jung, H., Mo, W., Green, H.W., 2009. Upper mantle seismic anisotropy resulting from
pressure-induced slip transition in olivine. Nature Geoscience 2, 7377.
Jung, H., Park, M., Jung, S., Lee, J., 2010. Lattice preferred orientation, water content, and
seismic anisotropy of orthopyroxene. Journal of Earth Sciences 21, 555568.
Karato, S., 1987. Scanning electron-microscope observation of dislocations in olivine.
Physics and Chemistry of Minerals 14, 245248.
Karato, S., 1995. Effect of water on seismic wave velocities in the upper mantle. Proceedings
of the Japan Academy. Series B, Physical and Biological Sciences 71, 6166.
Karato, S., 2003. Mapping water content in the upper mantle. Geophysical Monograph
138, 135152.
Karato, S., Jung, H., 2003. Effects of pressure on high-temperature dislocation creep in
olivine. Philosophical Magazine, A 83, 401414.
Karato, S.I., Toriumi, M., Fujii, T., 1980. Dynamic recrystallization of olivine single crystals during high-temperature creep. Geophysical Research Letters 7, 649652.
Karato, S., Paterson, M.S., Fitzgerald, J.D., 1986. Rheology of synthetic olivine aggregates: inuence of water and grain size. Journal of Geophysical Research 91,
81518176.
Karato, S.I., Jung, H., Katayama, I., Skemer, P., 2008. Geodynamic signicance of seismic
anisotropy of the upper mantle: new insights from laboratory studies. Annual Review of Earth and Planetary Sciences 36, 5995.

H. Jung et al. / Tectonophysics 594 (2013) 91102


Katayama, I., Karato, S.I., 2006. Effect of temperature on the B- to C-type fabric transition in olivine. Physics of the Earth and Planetary Interiors 157, 3345.
Katayama, I., Jung, H., Karato, S.I., 2004. New type of olivine fabric from deformation
experiments at modest water content and low stress. Geology 32, 10451048.
Katayama, I., Karato, S.I., Brandon, M., 2005. Evidence of high water content in the deep
upper mantle inferred from deformation microstructures. Geology 33, 613616.
Khisina, N.R., Wirth, R., Andrut, M., Ukhanov, A.V., 2001. Extrinsic and intrinsic mode of
hydrogen occurrence in natural olivines: FTIR and TEM investigation. Physics and
Chemistry of Minerals 28, 291301.
Kneller, E.A., van Keken, P.E., Karato, S., Park, J., 2005. B-type olivine fabric in the mantle
wedge: Insights from high-resolution non-Newtonian subduction zone models.
Earth and Planetary Science Letters 237, 781797.
Kohlstedt, D.L., Mackwell, S.J., 1998. Diffusion of hydrogen and intrinsic point defects in
olivine. Zeitschrift Fur Physikalische Chemie-International Journal of Research in
Physical Chemistry & Chemical Physics 207, 147162.
Kohlstedt, D.L., Goetze, C., Durham, W.B., Vander Sande, J., 1976. New technique for
decorating dislocations in olivine. Science 191, 10451046.
Kohlstedt, D.L., Keppler, H., Rubie, D.C., 1996. Solubility of water in the alpha, beta and
gamma phases of (Mg, Fe)(2)SiO4. Contributions to Mineralogy and Petrology 123,
345357.
Lewis, I.R., Chafn, N.C., Gunter, M.E., Grifths, P.R., 1996. Vibrational spectroscopic studies
of asbestos and comparison of suitability for remote analysis. Spectrochimica Acta
Part A: Molecular and Biomolecular 52, 315328.
Long, M.D., Silver, P.G., 2008. The subduction zone ow eld from seismic anisotropy: a
global view. Science 319, 315318.
Long, M.D., Silver, P.G., 2009. Shear wave splitting and mantle anisotropy: measurements, interpretations, and new directions. Surveys in Geophysics 30, 407461.
Lu, R., Keppler, H., 1997. Water solubility in pyrope to 100 kbar. Contributions to
Mineralogy and Petrology 129, 3542.
Mackwell, S.J., Kohlstedt, D.L., Paterson, M.S., 1985. The role of water in the deformation of olivine single crystals. Journal of Geophysical ResearchSolid Earth and
Planets 90, 13191333.
Mainprice, D., 1990. A fortran program to calculate seismic anisotropy from the lattice
preferred orientation of minerals. Computers & Geosciences 16, 385393.
Mainprice, D., 2007. Seismic anisotropy of the deep earth from a mineral and rock physics
perspective. In: Schubert, G. (Ed.), Treatise on Geophysics. Elsevier, Amsterdam,
pp. 437491.
Mainprice, D., Silver, P.G., 1993. Interpretation of SKS-waves using samples from the subcontinental lithosphere. Physics of the Earth and Planetary Interiors 78, 257280.
Mainprice, D., Tommasi, A., Couvy, H., Cordier, P., Frost, D.J., 2005. Pressure sensitivity of olivine slip systems and seismic anisotropy of Earth's upper mantle. Nature 433, 731733.
Manning, C.E., Menold, C.A., Yin, A., 2001. Metamorphism and exhumation of ultrahighpressure eclogites and gneisses, north Qaidam, China. Geological Society of America,
Abstracts with Programs 33 (6), 252.
Mei, S., Kohlstedt, D.L., 2000. Inuence of water on plastic deformation of olivine aggregates 1. Diffusion creep regime. Journal of Geophysical ResearchSolid Earth 105,
2145721469.
Mellini, M., Fuchs, Y., Viti, C., Lemaire, C., Linares, J., 2002. Insights into the antigorite
structure from Mssbauer and FTIR spectroscopies. European Journal of Mineralogy 14, 97104.
Menold, C.A., Manning, C.E., Yin, A., 2001. Metamorphism and exhumation of very high
pressure eclogites, North Qidam, China. EOS Transactions of the American Geophysical Union 82, 47.
Miller, G.H., Rossman, G.R., Harlow, G.E., 1987. The natural occurrence of hydroxide in
olivine. Physics and Chemistry of Minerals 14, 461472.
Mizukami, T., Wallis, S.R., Yamamoto, J., 2004. Natural examples of olivine lattice preferred
orientation patterns with a ow-normal a-axis maximum. Nature 427, 432436.
Mckel, J.R., 1969. Structural petrology of the garnetperidotite of Alpe Arami (Ticino,
Switzerland). Leidse Geologische Mededelingen 42, 61130.
Nakajima, J., Hasegawa, A., 2004. Shear-wave polarization anisotropy and subductioninduced ow in the mantle wedge of northeastern Japan. Earth and Planetary Science Letters 225, 365377.
Ohuchi, T., Kawazoe, T., Nishihara, Y., Nishiyama, N., Irifune, T., 2011. High pressure and
temperature fabric transitions in olivine and variations in upper mantle seismic
anisotropy. Earth and Planetary Science Letters 304, 5563.
Oneill, H.S.C., Wood, B.J., 1979. Experimental-study of FeMg partitioning between
garnet and its calibration as a geothermometer. Contributions to Mineralogy and
Petrology 70, 5970.
Panozzo, R.H., 1983. Two-dimensional analysis of shape-fabric using projections of
digitized lines in a plane. Tectonophysics 95, 279294.
Panozzo, R., 1984. Two-dimensional strain from the orientation of lines in a plane. Journal
of Structural Geology 6, 215221.
Park, J., Levin, V., 2002. Seismic anisotropy: tracing plate dynamics in the mantle. Science
296, 485489.
Park, M., Jung, H., Kil, Y., submitted for publication. Two contrasting petrofabrics of olivine in metasomatized spinel peridotites from the Rio Grande rift and their implications for seismic anisotropy. Contributions to Mineralogy and Petrology.
Paterson, M.S., 1982. The determination of hydroxyl by infrared absorption in quartz,
silicate glasses and similar materials. Bulletin de Mineralogie 105, 2029.
Petry, R., Mastalerz, R., Zahn, S., Mayerhofer, T.G., Volksch, G., Viereck-Gotte, L., KreherHartmann, B., Holz, L., Lankers, M., Popp, J., 2006. Asbestos mineral analysis by UV
Raman and energy-dispersive X-ray spectroscopy. Chemphyschem 7, 414420.
Post, J.L., Borer, L., 2000. High-resolution infrared spectra, physical properties, and micromorphology of serpentines. Applied Clay Science 16, 7385.
Prior, D.J., Boyle, A.P., Brenker, F., Cheadle, M.C., Day, A., Lopez, G., Peruzzo, L., Potts, G.J.,
Reddy, S., Spiess, R., Timms, N.E., Trimby, P., Wheeler, J., Zetterstrom, L., 1999. The

101

application of electron backscatter diffraction and orientation contrast imaging in


the SEM to textural problems in rocks. American Mineralogist 84, 17411759.
Raterron, P., Chen, J., Li, L., Weidner, D., Cordier, P., 2007. Pressure-induced slip-system
transition in forsterite: single-crystal rheological properties at mantle pressure and
temperature. American Mineralogist 92, 14361445.
Rinaudo, C., Belluso, E., Gastaldi, D., 2004. Assessment of the use of Raman spectroscopy
for the determination of amphibole asbestos. Mineralogical Magazine 68, 455465.
Ringwood, A.E., Major, A., 1971. Synthesis of majorite and other high-pressure garnets
and perovskites. Earth and Planetary Science Letters 12, 411418.
Russ, J.C., Dehoff, R.T., 2000. Practical Stereology, 2nd ed. Kluwer Acad, New York (381 pp.).
Shi, R.D., Grifn, W.L., O'Reilly, S.Y., Zhao, G.C., Huang, Q.S., Li, J., Xu, J.F., 2010. Evolution
of the Luliangshan garnet peridotites in the North Qaidam UHP belt, Northern
Tibetan Plateau: constraints from ReOs isotopes. Lithos 117, 307321.
Skemer, P., Katayama, I., Jiang, Z., Karato, S.-I., 2005. The misorientation index: development of a new method for calculating the strength of lattice-preferred orientation.
Tectonophysics 411, 157167.
Skemer, P., Katayama, I., Karato, S.I., 2006. Deformation fabrics of the Cima di Gagnone
peridotite massif, Central Alps, Switzerland: evidence of deformation at low temperatures in the presence of water. Contributions to Mineralogy and Petrology
152, 4351.
Skemer, P., Warren, J.M., Hirth, G., 2012. The inuence of deformation history on the
interpretation of seismic anisotropy. Geochemistry, Geophysics, Geosystems 13.
Skogby, H., Rossman, G.R., 1991. The intensity of amphibole OH bands in the infrared
absorption spectrum. Physics and Chemistry of Minerals 18, 6468.
Song, S.G., Yang, J.S., Xu, Z.Q., Liou, J.G., Shi, R.D., 2003. Metamorphic evolution of the
coesite-bearing ultrahigh-pressure terrane in the North Qaidam, Northern Tibet,
NW China. Journal of Metamorphic Geology 21, 631644.
Song, S.G., Zhang, L.F., Niu, Y.L., 2004. Ultra-deep origin of garnet peridotite from the North
Qaidam ultrahigh-pressure belt, Northern Tibetan Plateau, NW China. American
Mineralogist 89, 13301336.
Song, S.G., Zhang, L.F., Chen, J., Liou, J.G., Niu, Y.L., 2005a. Sodic amphibole exsolutions
in garnet from garnetperidotite, North Qaidam UHPM belt, NW China: Implications for ultradeep-origin and hydroxyl defects in mantle garnets. American
Mineralogist 90, 814820.
Song, S.G., Zhang, L.F., Niu, Y.L., Su, L., Jian, P., Liu, D.Y., 2005b. Geochronology of diamondbearing zircons from garnet peridotite in the North Qaidam UHPM belt, Northern
Tibetan Plateau: a record of complex histories from oceanic lithosphere subduction
to continental collision. Earth and Planetary Science Letters 234, 99118.
Song, S.G., Zhang, L.F., Niu, Y.L., Su, L., Song, B.A., Liu, D.Y., 2006. Evolution from oceanic
subduction to continental collision: a case study from the Northern Tibetan Plateau
based on geochemical and geochronological data. Journal of Petrology 47, 435455.
Song, S.G., Su, L., Niu, Y.L., Zhang, L.F., Zhang, G.B., 2007. Petrological and geochemical
constraints on the origin of garnet peridotite in the North Qaidam ultrahighpressure metamorphic belt, northwestern China. Lithos 96, 243265.
Song, S.G., Niu, Y.L., Zhang, L.F., Bucher, K., 2009. The Luliangshan garnet peridotite
massif of the North Qaidam UHPM belt, NW Chinaa review of its origin and metamorphic evolution. Journal of Metamorphic Geology 27, 621638.
Tasaka, M., Michibayashi, K., Mainprice, D., 2008. B-type olivine fabrics developed in
the fore-arc side of the mantle wedge along a subducting slab. Earth and Planetary
Science Letters 272, 747757.
Van Der Wal, D., Chopra, P., Drury, M., Fitz Gerald, J.D., 1993. Relationships between dynamically recrystallized grain-size and deformation conditions in experimentally
deformed olivine rocks. Geophysical Research Letters 20, 14791482.
Van Roermund, H.L.M., Drury, M.R., 1998. Ultra-high pressure (P > 6 GPa) garnet peridotites in Western Norway: exhumation of mantle rocks from >185 km depth.
Terra Nova 10, 295301.
Van Roermund, H.L.M., Drury, M.R., Barnhoorn, A., De Ronde, A.A., 2000. Super-silicic
garnet microstructures from an orogenic garnet peridotite, evidence for an ultradeep (>6 GPa) origin. Journal of Metamorphic Geology 18, 135147.
Wang, D.J., Mookherjee, M., Xu, Y.S., Karato, S., 2006. The effect of water on the electrical conductivity of olivine. Nature 443, 977980.
Wang, Q., Xia, Q.-K., O'Reilly, S.Y., Grifn, W.L., Beyer, E.E., Brueckner, H.K., 2013. Pressureand stress-induced fabric transition in olivine from peridotites in the Western Gneiss
Region (Norway): implications for mantle seismic anisotropy. Journal of Metamorphic Geology 31, 93111.
Withers, A.C., Wood, B.J., Carroll, M.R., 1998. The OH content of pyrope at high pressure.
Chemical Geology 147, 161171.
Xiong, Q., Zheng, J.P., Grifn, W.L., O'Reilly, S.Y., Zhao, J.H., 2011. Zircons in the
Shenglikou ultrahigh-pressure garnet peridotite massif and its country rocks
from the North Qaidam terrane (western China): Meso-Neoproterozoic crustmantle coupling and early Paleozoic convergent plate-margin processes. Precambrian Research 187, 3357.
Xu, Z.Q., Wang, Q., Ji, S.C., Chen, J., Zeng, L.S., Yang, J.S., Chen, F.Y., Liang, F.H., Wenk, H.R.,
2006. Petrofabrics and seismic properties of garnet peridotite from the UHP Sulu
terrane (China): implications for olivine deformation mechanism in a cold and
dry subducting continental slab. Tectonophysics 421, 111127.
Yang, J.J., Powell, R., 2008. Ultrahigh-pressure garnet peridotites from the
devolatilization of sea-oor hydrated ultramac rocks. Journal of Metamorphic Geology 26, 695716.
Yang, J.J., Godard, G., Kienast, J.R., Lu, Y.Z., Sun, J.X., 1993. Ultrahigh-pressure (60 kbar)
magnetite-bearing garnet peridotites from northeastern Jiangsu, China. Journal of
Geology 101, 541554.
Yang, J.S., Xu, Z.Q., Song, S.G., Wu, C.L., Shi, R.D., Zhang, J.X., Wan, Y.S., Li, H.B., Jin, X.C.,
Jolivet, M., 2000. Discovery of eclogite in Dulan, Qinghai Province and its signicance for studying the HPUHP metamorphic belt along the central orogenic belt
of China. Acta Geologica Sinica 74, 156168.

102

H. Jung et al. / Tectonophysics 594 (2013) 91102

Ye, K., Cong, B.L., Ye, D.I., 2000. The possible subduction of continental material to
depths greater than 200 km. Nature 407, 734736.
Yin, A., Rumelhart, P.E., Butler, R., Cowgill, E., Harrison, T.M., Foster, D.A., Ingersoll, R.V.,
Zhang, Q., Zhou, X.Q., Wang, X.F., Raza, A., 2002. Tectonic history of the Altyn Tagh
fault in northern Tibet inferred from Cenozoic sedimentation. Geological Society of
America Bulletin 114, 12571295.
Yin, A., Manning, C.E., Lovera, O., Menold, C.A., Chen, X., Gehrels, G.E., 2007. Early Paleozoic
tectonic and thermomechanical evolution of ultrahigh-pressure (UHP) metamorphic
rocks in the northern Tibetan Plateau, Northwest China. International Geology Review
49, 681716.

Yoshino, T., Matsuzaki, T., Yamashita, S., Katsura, T., 2006. Hydrous olivine unable to
account for conductivity anomaly at the top of the asthenosphere. Nature 443,
973976.
Zhang, R.Y., Liou, J.G., Yang, J.S., Yui, T.F., 2000. Petrochemical constraints for dual origin of
garnet peridotites from the DabieSulu UHP terrane, eastern-central China. Journal of
Metamorphic Geology 18, 149166.
Zhang, J.X., Yang, J.S., Mattinson, C.G., Xu, Z.Q., Meng, F.C., Shi, R.D., 2005. Two contrasting
eclogite cooling histories, North Qaidam HP/UHP terrane, western China: petrological
and isotopic constraints. Lithos 84, 5176.

Das könnte Ihnen auch gefallen